Sei sulla pagina 1di 388

The Rotary

Cement Kiln
Second Edition

bY
Kurt E. Peray

Edward Arnold
Preface
Often regarded as the heart of the plant, the kiln constitutes clearly the
most important step in the process of cement manufacturing. It represents
the largest single capital investment and consumes the major portion of
the energy requirements in the plant. Regardless how much effort and
attention is being given to the preparation of the kiln feed, the fact remains
that the feed has to be properly burned in the kiln so that a good quality
product can be sold to the customer. Because of its importance, the kiln
burning operation deserves special attention and kiln operators should be
properly trained. The old simple saying still holds true: “When the kiln
discharges clinker, the company has a fighting chance to make some
profits, but when no clinker is produced, no money can be made.”
The rotary kiln requires specialized knowledge and experience on the
part of the operator so he can successfully perform his job. Thus, with its
complex instrumentation and multiple reactions, the kiln poses a
significant challenge to the kiln operator. It is obvious that the kiln
operator occupies one of the key positions in the production crew.
2% Rotary Cement Kiln is the first handbook of its kind to deal not
only with the theoretical aspect, but also with the actual control functions
of kiln operation. First .published in 1972, the original edition of this
book dealt primarily with wet- and long dry-process kilns. Since that
time, the cement industry has undergone a radical change brought about by
the energy crisis of the mid-seventies. More fuel and labor efficient kilns
were built to keep pace with the rapid advances in cement manufacturing
technology. Capital that was sufficient twenty years ago to buy a
complete new cement plant today will barely be enough to buy a kiln.
But, the new modem preheater and precalciner kilns of today, outperform
and outproduce the older wet and dry kilns by a wide margin. Most of
these are also fully automatic, controlled by computers and the noisy,
dusty burnerfloor of the past has been replaced by remote, air-conditioned
control rooms. There is no question that these technological advances
have benefited the kiln operator for they have made his job easier and more
pleasurable. On the other hand, there is no doubt that the operator’s
responsibility has greatly increased because most are responsible not only
for kiln operation but for the control of the raw and finished grinding
departments too. It is the author’s hope that this revised edition will be as
well accepted in the cement industry as the first book. We have expanded
each chapter to make this a more complete and up-todate training and
reference book not only for kiln operators but for supervisors and
management staff as well. Most important, we have added extensive
discussions for preheater and pmcalciner operations.
The author discusses the theoretical fundamentals, including basic
cement chemistry, composition of the kiln feed, heat balances and heat
transfer, combustion, flames, fuels, and the air circuitry in a rotary kiln.
Step-by-step descriptions of the control functions for the operation of a
rotary kiln are extensively discussed The described burning procedures and
techniques have been tested over many years on kilns of various
dimensions, and experience has proven them to be entirely successful. So
much so that computer control programs have been recently written and
successfully placed in operation that were based on the 27 basic kiln
control conditions first introduced by the author in our first book. Adopted
for hundreds of kilns worldwide, they are the foundation for stable and
economical operations.
The appendix includes a section with conversion tables, definitions of
common terms relating to rotary kilns, and a suggested outline for a
training program for new operatots.
Many thanks to Joseph J. Waddell who coauthored the first edition with
me.
Contents
PART I. KILN SYSTEMS AND THEORY
1. History 3
2. Types of Rotary Kilns 6
3. The Refractory 17
4. Fuels 37
5. Combustion 44
6. The Flame 63
7. Heat Transfer 83
8. Heat Balances 106
9. The Chemistry of Kiln Feed and Clinker 115
10. Reaction Zones in the Rotary Kiln 141
11. Coating and Ring Formation in a Rotary Kiln 147
12. The Air Circuit in a Rotary Kiln 155
13. I’$ovement of Material Through the Kiln 174

PART II. KILN OPERATING PROCEDURES 199


1 4 . Kiln Operating and Control Methods 201
15. Instrumentation 207
1 6 . Kiln Control Variables 232
17. Fuel Systems 253
1 8 . Clinker Cooler Control 266
1 9 . Kiln Exit-Gas Temperature Control 292
20. Feed-Rate Control 299
2 1 . Kiln Starts and Shutdowns 308
22. The 27 Basic Kiln Conditions 327
2 3 . Kiln Emergency Conditions 346
24. Safety and Accident Prevention 362

Appendix A: The International System of Units (SI) 367


Appendix B: Weights and Measures 369
Appendix C: Temperature Conversions 378
Appendix D: Kiln Operator’s Quiz 382
Index 386
Part Z

Kiln Systems and Theory


1.
History
Vertical furnaces and simple forms of shaft kilns were used for burning
lime well over 2,000 years ago. History tells us that the Romans used a
vertical furnace in which to burn a pozzolanic lime. Near Riverside,
California are the remains of underground furnaces (Fig. 1.1) in which the
early Mexican settlers burned limestone to make lime during the fmt part
of the 19th century. In later times so-called bottle and shaft kilns were em-
ployed. Vertical kilns of the type shown in Fig. 1.2 were constructed in
Southern California about the turn of the century.
Early development of the rotary kiln probably started about 1877 in
England, but Frederick Ransome is usually credited with the fmt
successful rotary kiln, which he patented in England in 1885. Although
the fmt Ransome kilns were a major breakthrough in the cement industry
at that time, many years passed before a successfully operating rotary kiln
was put into production. It was mainly the pioneer work of American
engineers a few years after Ransome’s discovery that brought the concept
of the rotary kiln out of its infancy. The first economical rotary kiln in
America, developed by Hurry and Seaman of the Atlas Cement Company,
went into production in 1895.
Shaft kilns with continuous feed are now used mainly and only for the
burning of lime and minerals other than cement. Rotary kilns have re-
placed these shaft kilns entirely in the cement industry. Although years
ago, shaft kilns showed lower thermal and power requirements than rotary
kilns, the advent of the preheater and precalciner kilns with their increased
output and fuel efficiency has apparently made the shaft kiln obsolete for
the burning of cement clinker.
THE ROTARY CEMENT KILN

Fig. I . 1 Remains of underground furnaces that were used by early


California Settlers for burning limestone to make lime. (Riverside Division,
American Cement Corp.)

The first Ransome kilns were 45 cm (18 in.) in diameter and 4.5 m
(15 ft) in length. Later, about 1900, the rotary kiln grew to 1.8 m (6 ft)
in diameter by 18 m (60 ft) long which in todays terms would have to be
classified as miniatures. Kiln sizes really started to explode in the 1960’s
when they reached dimensions up to 6.5 m (21 ft) diameter and up to
238 m (780 ft) length. With these enormous sizes and corresponding high
output rates a considerable amount of new structural and control problems
started to evolve. Refractory life in the kiln became uneconomically low,
coolers couldn’t handle that high output especially not during upset condi-
tions, and mechanical equipment failures became weekly occurrences in
many plants.
The energy crisis represented a blessing in disguise in matters of kiln
design. Suddenly, fuel conservation became the number one priority item
in most cement plants which led directly to increased construction of
preheater kilns all over the North American continent. Although these pre-
HISTORY

Fig.12 Vertical shaft kilns were commonly in use in the latter pad of the
19th century. (Riverside Division, American Cement Corp.)

heater kilns satisfied the need for lower fuel consumption, they didn’t meet
tbe requirements for using low-grade fuel and ever-increasing demands for
higher production rates.
In an attempt to gain these higher outputs, the Japanese cement indus-
try increased preheater kiln sizes to a point where they were back to square
one, namely, these kilns again became too large; frequent mechanical
problems and short brick-life became the norm just as in the times of the
dry and wet monster kilns. The major breakthrough came in Europe where
precalcination was successfully attempted in the late 1960’s using a very
low bituminous shale as a component of the kiln feed in a conventional
preheater kiln. Adding combustible materials to the kiln feed, at that time,
was nothing revolutionary, for the author himself, in 1957, had burned a
wet kiln in Canada that contained oil shale in the slurry. The European
experience, however, was the first time such an addition was successfully
tried in a preheater kiln and thus paved the way for today’s precalciner kiln.
Precalciner kilns are the latest advance in cement manufacturing
technology. They combine low thermal requirements, are able to use low-
grade fossil fuels or other combustible materials, and show output rates
that were considered unattainable only a few years back.
2.

Types of Rotary Kilns

Generally speaking, the clinker manufacturing processes used in rotary


kilns are classified into:
Wet-Process Kilns
Semidry Kilns
Dry-
Preheater Kilns
Precalciner Kilns

Each of these types am discussed here.

2.1 WET PROCESS

Into this group fall all processes in which the kiln feed enters the kiln
in the form of a slurry with a moisture content of 30 to 40%. In
comparison with a dry-process kiln of the same diameter, a wet-process
kiln needs an additional zone (dehydration zone) to drive off the water from
the kiln feed. Therefore, it must be considerably longer in or&r to achieve
the same production rate.
To produce an equivalent amount of clinker, a wet-process kiln requires
theoretically more fuel than a dry-process kiln because of the extra heat
required to evaporate the water. I-Iowever, in actual operation of a kiln this
fundamental fact does not always hold entirely true. As one progresses in
the reading of this book, the reasons for these discrepancies between theory
TYPES OF ROTARY KILNS

and actual operation will become clearer and understandable.


Advantages of a wet-process kiln are:

1. feed is blended more uniformly than in the dry process


2. dust losses are usually smaller, and
3. in moist climate regions, wet processing of the raw material is
more suitable than dry because of moisture already present in the
blend materials.

2.2 SEMIDRY PROCESS

This member in the group of rotary kilns is also widely known under
the term Grate Process Kiln or L.qol Kiln. These kilns axe as efficient in
matters of fuel consumption as the most modern preheater and pmcalciner
kilns. Output rates, however, lag behind the aforementioned types of
kilns. However, it is advantageous to select a Grate Process Kiln over a
preheater or precalciner kiln in places where raw material moisture is so
high that it cannot be economically dried by waste heat from the kiln.
Lepol Kilns, because of the fact that the kiln exit gases pass through the
granular feed bed, operate with much lower dust contents in the waste
gases which gives these kilns a decisive advantage over other preheater
kilns. Instead of granulating the kiln feed, some plants use filter press
cakes to feed the kiln. In such cases, the wet-kiln feed slurry is first passed
through large presses for removal of the free water and more importantly,
to remove alkalies before the filter cakes are fed to the kiln.
In the grate process, pulverized dry-kiln feed is first pellet&d into small
nodules by means of 10-U% water addition, then the nodules are fed onto
a traveling grate where they are partly calcined before they enter the rotary
kiln. Heating of the nodules is effected by the exit gases from the rotary
kiln, the hot gases passing through the material bed from above as they are
drawn downward through the grates by means of a fan. The partly calcined
material then falls down a chute into the rotary kiln where final clinkeriza-
tion takes place. Because the kiln feed is already partly calcined before it
enters the kiln, the rotary kiln itself is only about one-third the usual
length. Fig. 2.1 is a schematic diagram of the flow of gas and material
through a Lepol grate-process preheater.
THE ROTARY CEMENT KILN

- - - - MATERIAL FLOW
SAS FLOW

Fig. 2.1 Flow diagram of a Lepol preheater. Pelletized feed fed onto a
traveling grate. is heated and ptly calcined by hot kiln exit gases before it
enters the kiln.

One advantage of grate-process kilns is the uniform size of clinker


leaving the kiln, an aspect that is decidedly beneficial for grinding the
clinker. However, there are some features not found in conventional rotary
kilns that need very close attention; for example, production of the nodules
and control of the thickness of the feed bed over the traveling grates. Such
a kiln usually requires additional labor to attend the granulator plant

2.3 DRY-PROCESS KILNS

As the term indicates, in this process the kiln feed enters the kiln in
dry powder form. Dry-process kiln dimensions are similar to wet kilns in
that they are long and typically show a length-to-diameter ratio of approxi-
mately 30: 1 to 35: 1. Dry-process kilns operate with a very high, back-end
temperature and require watersprays at the feed end to cool the exit gases to
safe levels before they enter the baghouse or precipitator. Most dry kilns
are equipped with chain sections at the feed end to transfer heat, that other-
wise would be lost, to the feed before the gases leave the kiln.
Fig. 2.2 shows a picture of a chain section. The gases enter the chains
at a temperature of approximately 800 C (1470 F) and leave the kiln exit
at a temperature of 450 C (840 F). In countercurrent flow, the material
TYPES OF ROTARY KILNS

enters the chains with a temperature of 50 C (120 F) and emerges from


the chain section with a temperature of 730 C (1350 F). Chain sections
are a high maintenance item; difficult to repair but an absolute necessity
for efficient operation. Because of the high costs of these chains, there is a
tendency in many plants to neglect proper and frequent maintenance. How-
ever, it has been found in many instances that the costs saved, by not tak-
ing care of the chain system, are coming back manyfold in the form of
higher fuel operating costs.

Fig. 2.2 The chain section of a kiln. ?he chains absorb heat fmm the hot
gas and transmit tbe heat to tbe kiln feed.

There is an advantage found in dry kilns that none of the other kiln
types exhibit. The high exit-gas temperature on these kilns renders them
perfectly suitable for cogeneration of electrical power. As a matter of fact
there are several plants with dry kilns that generate their own power and
many existing plants are taking a hard look at the feasibility of adding a
power plant to their facility. The reasoning is simply that generating
power is energy-conserving and in some locations it may be more eco-
nomical to add a power plant to an existing dry kiln than to convert this
kiln to preheater status.
THE ROTARY CEMENT KILN

2.4 THE PREHEATER KILN

In the gas-suspension preheater kiln, the dry feed is preheated and partly
calcined by the hot kiln exit gases in a tower of heat exchange cyclones.
This concept is, contrary to popular beliefs, not new because a patent on
this type of kiln was issued in Czechoslovakia already in the early 1930’s.
However, the suspension preheater kiln as it is known today did not come
into its own right until after World War II when German kiln manufac-
turers were able to overcome the operational and structural problems of
these types of kilns (Fig. 2.3).
The preheating of the kiln feed is done outside the rotary kiln proper,
i.e., before the feed enters the kiln. The heat exchange between the gas and
the material takes place in the cyclones while both are in suspension.
Many different designs of preheater towers are in existence using this basic
principle. The most common design is the parallel four-stage preheater.
Some of these can reach output rates of up to 8000 metric tons per day.
Exit gas temperatures leaving the top # 4 stage are around 340 C (640 F)
and in many such plants, these waste gases are used (together in some
cases with the waste gases from the clinker cooler) for drying and
preheating of kiln feed in the raw grinding department. One drawback of
preheater kilns is the high concentration of volatile constituents such as
alkalies, sulfur, and chlorides in the kiln exit gases that give rise to
numerous plug-up problems at the lower cyclone stage and kiln inlet. For
this reason, most suspension preheater kilns must be equipped with an
alkali and sulfur bypass that allows evacuation of a percentage of the kiln
exit gases and thus bypasses the preheater cyclones. Such bypasses are not
only used for reducing plug-up problems but in many plants are a neces-
sity to keep the alkali content in the clinker below maximum permissible
levels.
Suspension preheater kilns are the most energy-efficient types of kilns
available operating typically with a specific fuel consumption of around
3138 MJ/ton clinker (750 k&kg, 2.7 MBtu/sh.ton).

2.5 THE PRECALCINER KILN

Roughly 15 years ago, Japanese cement manufacturers were confronted


with the question of how best to increase the production rates of existing
preheater kilns. As mentioned earlier, their preheater kilns reached limits
TYPES OF ROTARY KILNS

FEED

- - - MATERIAL FLOW
- SAS F L O W

KILN
Fig. 2.3 Flow diagram of a Humboldt gas su.spension preheater, a multistage
counteiflow process which draws the gases through a series of cyclone
collectors. The pulverized material, which is fed counter to tbe gas flow,
becomes suspended in the gas stream and is heated in successive stages until
discharged through the kiln-feed pipe.
THE ROTARY CEMENT KILN

in dimensions that gave rise to considerable operational and structural


design problems. These kiln diameters became so large that refractory ser-
vice life problems caused frequent and costly kiln down times. Then there
was also the question of the possibility of using low-grade fuels in cement
kilns that had to be addressed. Out of these considerations evolved the con-
cept of auxiliary firing to precalcine the feed outside the rotary kiln. FVe-
calciners are essentially suspension preheater kilns that are equipped with a
secondary firing system (commonly referred to as flash furnaces) attached
to the lower stage of the preheater tower. This allowed kiln manufacturers
to construct smaller diameter kilns without sacrificing kiln output, for a
pmcalciner produces 50-70% more clinker than a conventional preheater
kiln of equal diameter. A precalciner, however, does not operate at a lower
specific total fuel consumption than a preheater kiln; their consumptions
are approximately the same. But by burning 30-50% of the total energy
input at the rear of the kiln, the heat load in the burning zone proper is
reduced producing a beneficial effect on refractory service life. Aside from
this, being able to use low-grade cheaper fuels in the auxiliary firing unit
results in a reduction of the costs per unit weight of clinker. Therefore, al-
though the energy consumption is the same, the cost for this energy is
usually lower-a decisive factor in favor of the precalciner kiln.
There are two different types of precalciner kilns-kilns with tertiary air
ducts and kilns without. As is discussed later in more detail, combustion
of any type of fuel requires a given amount of air, hence this air has to be
provided to the auxiliary firing unit at the back end of the kiln. This is
done in two ways, either, the air comes from the kiln itself (no tertiary
duct) or it is supplied by the enormous excess waste gases from the clinker
cooler by means of the tertiary air duct that runs parallel to the kiln. he-
calciners without tertiary air ducts can be equipped with any kind of a
cooler including planetary coolers whereas kilns with tertiary ducts cannot
have planetary coolers. Kilns with tertiary ducts are more difficult to
control for the operator since these kilns contain two distinct and separate
combustion processes that must be closely controlled independently. How-
ever, the precalciner without the cooler air duct tends to be less efficient in
fuel economy than the other when a large percentage of the kiln exit gases
have to be bypassed in the preheater tower. Installation of a tertiary air
duct is also very expensive and constitutes a high maintenance item
In the short time span of 90 years, rotary kilns have undergone vast
changes. Great improvements in fuel efficiency and control technology
have been made and most newly constructed kilns are fully automatic and
corntolled by computers. With a rating of 55%, the modern efficient
TYPES OF ROTARY KILNS

\
GAS OUT-

FE,,,,2 1,
1. STAGE

2. STAGE

3. STAGE

OIL BURNER

4. STAGE FLASH FURNACE

COOLER

Fig. 2.4 SF-Suspension flash preheater kiln.

cement kilns appear to have approached the limit of best attainable fuel
efficiency and output, Have kiIn manufacturers reached optimum kiln de-
sign? Is the precalciner kiln the kiln of the future? The author and many
cement engineers do not think so. Technology has never stood still in an
age of rapidly accelerating industrial changes. Cement kilns will not be
exempted from this trend. Although the precalciner kiln is an important
piece of process equipment, it has numerous shortcomings such as the
enormity of the tall tower on the kiln backend, and problematic
THE ROTARY CEMENT KILN

1. STAGE

2. STAGE

3. STAGE

4. STAGE

\
KILN

Fig. 2.5 KSV-!$outed bed and votix chamber precalciner kiln.

environmental conditions. Because these systems have become so


sophisticated in matters of control and maintenance, well-trained specialists
are now an absolute requirement. There are many new concepts for
burning cement clinker that have been advanced by a multitude of forward-
thinking engineers. The fluid bed reactor, the separation of limestone
calcining from other raw material preheating, the two-stage traveling-grate
preheater am just some of the more “exotic” ideas that have come to the
forefront. Perhaps one of these will be the kiln of the future, so it is
important to keep an open mind toward such ideas and not regard them as
“not possible.”
GAS

1. STAGE

2. STAGE

3 . STAGE

TERTIARY DUCT
4. STAGE
FROM CDDLER

:
\
\

Fig. 2.6 RSP-Reinforced suspensions preheater kiln.


THE ROTARY CEMENT KILN

1. STAGE

2. STAGE

3. STAGE

4. STAGE

MFC
FLUIDIZED
CALCINATCR

KILN

BLOWER
ER
PRECIPITATOR

Fig. 2.7 MFC-Fluidized calcinator kiln.


3.

The Refractory

Because of the high temperatures existing inside a kiln during the


clinker manufacturing process, it is necessary to protect the steel shell of
the kiln with a refractory lining. If this protection were not provided the
shell would disintegrate within a few hours. A refractory is a material,
usually nonmetallic, that is used to withstand high temperature. In a kiln,
the refractory usually consists of brick of special composition and sizes as
shown in Fig. 3.1. Some usage in recent years has been made of a cast
lining continuously placed in a manner similar to placing concrete in a
structure. In this method, the interior is progressively formed by means of
special planks, welded anchors, and snapties. The kiln is rotated as
necessary during placement of each section of lining so the workmen are
always working at the same level.
Among kiln operators, refractory failure is considered the most critical
upset in a kiln operation. Refractory failure inside the rotary kiln is
indicated when the kiln shell becomes red hot because the refractory lining
has either been entirely lost or has become so thin in an area that the kiln
shell becomes overheated Such a condition is dangerous because once the
protection supplied by the refractory has been removed, the steel shell
could easily be warped to such an extent that replacement of an entire kiln
shell section becomes necessary. In most instances, however, damage can
be avoided if the kiln is shut down for lining replacement as soon as the
shell starts to show a large red spot. Because of the importance of such a
situation, remedial procedures for hot shell conditions are described in
Chapter 25.
Replacement of the kiln lining, especially in the burning zone, is
unfortunately a frequent necessity, exerting a large strain on the operating
9X4’/4%9 Arch
9X4’12X4 9x6x3
Calliau block Wedge
cupola block BASIC
CLAY

9x6x3’/2
Arch 9x6x3’/2
Wedge
ROTARY KILN BLOCKS

RKB

Fig. 3.1 Bricks for the nzhctory lining come in several sizes and shapes
(Kaiser Rcfacfories.).
THE REFRACTORY

budget and on production schedules. For example, replacement of an entire


burning zone lining over a length of 50 ft in a 16-ft diameter kiln costs in
excess of $80,000. This figure represents only the cost of the refractory
itself and does not include the added expense of installation, nor the hidden
cost resulting from loss of production and the extra fuel required to bring
the kiln back to operating temperatures.
Ideally, one would like to obtain a service life on the burning-zone
lining that would extend through a continuous operation cycle from one
turn-around to the next. In other words, the optimum service life would be
a period of either 11 or 23 months. In most kilns, however, this is more
the exception than the rule. Although linings in the upper, cooler regions
of the kiln can show lives of 5-20 years, in the intensely hot burning zone
the life ranges from as little as 30 days to as much as two years.
Regardless of the fact that all cement kilns operate within a narrow
temperature range in the burning zone there still exists that large
discrepancy between kilns in matters of refractory service life. The reason
for this can be found in the fact that each kiln has its own specific
characteristics and idiosynchrasies which greatly influence service life. The
life of refractories in a kiln is governed by key factors which are discussed
in the following section.

3.1 FACTORS AFFECTING WEAR OF


REFRACTORIES
Frequency of Kiln Shutdowns.

Many plants have found that refractory life is often directly proportional
to the number of kiln shutdowns that were experienced while the refractory
was in the kiln. The more shutdowns and kiln stops, the shorter the life.
The danger of damaging the refractory is directly related to the rate of cool-
down of the kiln, the danger being the greatest when cooling is too rapid.
The first step in preventing this situation is obviously to eliminate shut-
downs by operating the kiln more efficiently on a continuous basis.
The second step is to make sure that cooling is slow and uniform when
the kiln is shut down. Cooling time should be at least 8 h and preferably
longer. A large guillotine damper to seal the kiln exit (back-end) helps to
conserve heat inside the kiln and retards cooling during a shutdown. On
some rotary kilns, retaining a small fire during the shutdown accomplishes
more or less the same results. Kilns equipped with internal heat
THE ROTARY CEMENT KILN

exchangers such as chains and crosses, however, are exempt from this
procedure, because leaving a fue in such a kiln under these conditions
might lead to chain failure due to the presence of oxygen at a fairly high
temperature without any cold feed entering the kiln. To secure a uniform
cooling, the kiln has to be turned and jacked on a regular schedule, because
the feed bed and the refractory underneath it will cool much slower than the
refractory which is exposed to the kiln gases. A suggested schedule for
turning the kiln after a shutdown on both dry and wet process kilns of
various lengths and diameters, given in Table 3.1, should ensure a uniform
cooling of the refractory and kiln shell (the kiln shell contracts also during
cooling).

TABLE 3.1
TYPICAL KILN JACKING SCHEDULE
lkqtency of tums should be as follows:
1. Continuous rotation on auxiliary drive for 30 min.
2 . One-third turnevery
10 min for 1 h.
3 . Onethird turn every
1 5 min f o r 1 h .
4 . Onehalf turn every
3 0 min f o r 4 h .
5. Onehalf turn every hour for 4 h.
6 . Onehalf turnevery
2 h for 12 h.
7. Onehalf turn once every 24 h regardless of length of shutdown.

NOTES: Each time before the kiln is turned, ascertain that no one is inside the kiln
and first cooler stage. Once a shift, check the clinker loading in the cooler inlet and
mn this material out when the pile is higher than 2 ft. This is necessary because each
time the kiln is turned a small amcunt of material is dumped into the cooler. During
periods of heavy rainfall it may be necessary to jack the kiln more frequently to
provide even cooling. This is a typical schedule for one paticular kiln. Actual plant
conditions determine the schedule for any other kiln.

Another procedure to secure slow cooling of the kiln is to make sure


the draft fan (I.D. fan) is shut down immediately when the fire has been
taken out of the kiln (an absolute must on a kiln with internal heat
exchangers). The primary air fan should be left running only for such a
time as needed to protect the burner pipe from heat during the early period
of kihl cooling.
THE RBFRACTORY

Too Rapid Heating of the Refractory.


Just as in rapid cooling, so heating the refractory too fast, can cause
thermal deterioration of the brick. The question of how much time should
be allowed to raise the kiln to normal operating temperature is subject to
wide differences of opinion. Recommendations by refractory manufacturers
vary to some extent depending on the type of refractory installed. One
factor sometimes overlooked is the kiln shell itself, which expands upon
heating. Governed partly by the thermal conductivity of the refractory,
this shell expansion takes place somewhat slower than expansion of the
bricks. Because of this, a period of at least 16 h should elapse when heat-
ing a kiln to operating temperature. This is on the conservative side.
After all, there is far less danger of damage when the temperature is raised
too slowly than when it is raised too fast. More time should be taken to
bring the kiln on line with thicker linings and larger kilns.

Overheating the Kiln.


One of the prime duties of the kiln operator is to make sure the kiln
never becomes dangerously overheated When a kiln is overheated the feed
starts to ball up and the coating turns to liquid in the burning zone. In a
matter of minutes, unless this condition is not drastically counteracted, the
entire refractory lining could be melted and lost as soon as the protective
coating has melted away. Overheated conditions are more apt to result
from erratic feed loading and advancement in the kiln rather than from any
action of the kiln operator. Indirectly, cycling and upset kiln operating
conditions can therefore be a cause for short refractory life. This then is
another reason why kiln operating stability is so important

Quality and Unifbrmity in Size and Shape of the


Refractory.
Since the costs of the refractories are low when compared to the costs
for the loss in kiln production, labor expenses, and fuel needed to reheat
the kiln, it is advisable to consider refractory quality and performance ahead
of the price of the refractory itself. On large diameter kilns it becomes
especially important that the refractory shapes do not deviate by more than
2 mm (0.078 in.) on each of its planes. Kiln managers also have to select
the right type of refractories for each location of the kiln which is easier
said than done. Experienced refractory salesmen that have a working
knowledge of rotary kilns can be of great help in this selection process.
THE ROTARY CEMENT KILN

Chemical Composition and Uniformity of the Kiln


Feed.

A basic lining in the burning zone must have a good protective coating
in or&r to achieve optimal service life. The type of coating is dependent
largely on the chemical composition of the feed and its uniformity.
Volatile constituents such as alkalies, sulfur, and chlorides can attack and
weaken a refractory lining. In short, a plant chemist must not only con-
cern himself with the ultimate cement quality a kiln feed will deliver but
must also design the mix to possess good bumability and coatability
properties in the kiln. This subject is discussed in more detail in the
section on kiln-feed chemistry.

Mech&cal Condition of the Tire and Kiln Shell.


Each kiln, with every revolution, undergoes shell deformations in the
vicinity of the tire that have a detrimental effect on the lining. This
deformation is generally referred to as shell ovality. There is no kiln that
is perfectly round. Ovality can be checked with the “HOLDERBANK”
shell test unit. Each plant should have a schedule to frequently check the
slippage of the tire at least twice a month as there is a direct relationship
between this slip and the shell ovality. Excessive slippage is an early
warning signal that shell ovality might be too high. Of equal importance
are annual checks of kiln alignment preferably when the kiln is in opera-
tion because misaligned kilns too, can result in excessive stresses on the
lining.

Poor Location and Directed Flame Patterns.

Effect of these are extensively discussed in the chapter on flame control.

Installation Methods of Refractories.

Quality of workmanship in refractory installations has a profound


influence on how well these refractories perform in a rotary kiln. A refrac-
tory liner can not achieve optimum life potential unless it has been
properly installed in accordance with time-proven methods and procedures.
Extra time and efforts exerted during construction of a lining can pay divi-
dends in longer service life and kiln uptime.
THE REFRACTORY

Proper Kiln Operating Procedures.

This includes the whole gambit of operating procedures discussed in


this book from combustion to burning zone control to kiln start and stop
procedures. Again, applying time-proven procedures and being consistent
in control follow-up can pay dividends in longer lining life. There .are no
short-cuts in this respect.

3.2 REFRACTORY REQUIREMENTS AND


PROPERTIES

Certain requirements have to be imposed upon the refractory, depending


on the conditions it will be exposed to in the area of the kiln where it is to
beused.

Resistance to High Temperatures.

The refractory has to withstand the temperatures that can prevail under
adverse conditions as well as those that prevail under normal conditions in
the zone where it is being used. Not only does it require the ability to
withstand high temperatures without melting, it also must maintain its
structural strength at temperatures below its melting point, and must main-
tain a constant volume when exposed for prolonged times to the high
temperatures.

Spalling Resistance

Any kiln shutdown, start-up, or severe operating upset usually creates


large temperature changes in the kiln, and the refractory must possess the
necessary shock resistance to withstand such temperature variations. Fail-
ure to possess this quality can cause the brick to crack. These cracks, gen-
erally referred to as spalling, develop in a horizontal direction on the brick
Spalling results from thermal shock. The same reaction can be
observed with a drinking glass: If we set a cold glass into very hot water
the thermal shock causes the glass to crack. However, if we slowly raise
the temperature of the water containing the glass, there is no thermal
shock and the glass is not damaged. The same applies to the refractory in
THE ROTARY CEMENT KILN

the kiln. When a cold kiln is fired, the temperature of the refractory must
be raised very slowly to avoid spalling.
Resistance to Chemical Attack (Slag Resistance).
During the process of clinkerizing, ash, slag, and vapors formed during
the combustion process can attack the refractory, reacting chemically with
brick, depending on the type of fuel used. Furthermore, dust and alkalies
entrained in the kiln gases can adhere to the bricks in the burning zone and
react with the refractory. The ability of a refractory to withstand these
chemical attacks is called its slag resistance. A brick that does not possess
this resistance could be considerably weakened by chemical attack,
resulting in premature refractory failure.
Abrasion Resistance.
Conditions encountered in a rotary kiln make it necessary that the
refractory withstand the abrasive action resulting from the sliding kiln feed
bed and also by dust entrained in the moving gas stream. This abrasion
resistance is a prerequisite for all bricks installed in front of and behind the
burning zone where coating is not usually form4

Coatabili@.
One of the most important qualities required from the refractory in the
burning zone where the highest temperatures exist, is its ability to take on
a good coating and to hold this coating for a prolonged length of time.
The importance of coating in the burning zone is discussed in Chapter 6.
Just as the refractory acts as a protection for the kiln-shell, so the coating
in turn acts as a protection for the refractory, thus serving to prolong the
life of the brick in the burning zone.

3.3 TESTS OF REFRACTORIES

Selection of the best refractory for a given area in the kiln is not an
easy task for the person who has to make this choice. This is especially
true for a kiln in a new cement plant where no previous experience can be
used as basis for this choice. To arrive at the proper selection, one has to
deal with a multitude of refractory manufacturers and literally hundreds of
different refractory types and shapes. Any one type of refractory can
THE REFRACTORY

perfotm extremely well in one kiln and be almost useless in another.


Because all kilns are different and many different raw materials are used to
produce clinker, it is essential that the refractory is designed specifically for
a particular kiln and its specific operating conditions.
Manufacturers of refractories usually furnish conventional information
on their materials, such as compressive and tensile strength, modulus of
elasticity, chemical analysis, thermal conductivity, density, porosity, and
gas permeability. In addition, there are various special properties, deter-
mined by certain tests that have become standardized in the refractory
industry. Results obtained from these tests, while not 100% conclusive,
do furnish a good indication of the properties of the refractory and its re;
sistance to various exposures within the kiln, and are the basis for the
selection of a refractory particularly suited to any given ama of the kiln.

Melting Point.
The melting point is the temperature at which the refractory starts to
sag and lose its structure or shape. However, the given melting point by
no means guarantees that the refractory will not start to sag at a lower
temperature. Depending on whether a reducing or an oxidization atmos-
phere exists in the area where the refractory is used, and also depending on
the composition of the refractory itself, melting can take place at a
temperature below the theoretical melting point. Bricks in prolonged use
and which have chemically reacted can have a melting point lower than
indicated by the test,
For a fire clay refractory, a test is used (ASTM C24) in which the PCE
(pyrometric cone equivalent) value is determined. PCE is the number of
the standard pyrometric cone whose tip would touch the supporting plaque
simultaneously with a cone of the refrctory material being investigated.
Together with this value, the manufacturer usually states the corresponding
melting point, which is the temperature at which the tested cone has
softened and sagged to the same extent as a standard cone. Here again one
has to remember that the maximum operating temperature should be well
below the melting point for reasons previously stated.

Hot Load (Temperature of Deformation).


Certain refractories, when under load (when subjected to a given static
pressure) undergo softening at a temperature far below the melting point.
Softening of the bricks can take place because of their own weight or
THE ROTARY CEMENT KILN

because of the pressure exerted upon them by the material bed in the rotary
kiln. Overheating of refractory can also cause deformation and thus reduce
the bearing strength under load. The test (ASTM C16) determines the
resistance to deformation or shear of refractory brick when subjected to a
specified compressive load at a certain temperature for a specified time.
The results of the hot load are expressed as the temperature at which a
definite deformation takes place when the brick is subjected to a given
static pressure, or the percent deformation at a stated temperature when
subjected to a given static load (usually 2 kg per cmZ or 25 lb per in?).
Because this test is performed over a relatively short period of time, it does
not give an indication of the structural strength of the refractory when
subjected to prolonged and constant high temperatures and loads. Such a
prolonged exposure could weaken the refractory further.

Linear Expansion or Shrinkage (Reheat Test).

As stated above, a refractory, under prolonged service at high


temperature and repeated heating, can undergo expansion or shrinkage. In
the reheat test, these changes in volume are determined and generally
referred to as irreversible (permanent) changes. The results am expressed in
percent linear change at a stated temperature, a plus (+) sign indicating
growth and a minus (-) sign, shrinkage. The test (ASTM C113) is con-
ducted by placing the refractory specimens in a down-draft kiln for several
hours, then cooling for at least 10 h, with length measurements made
before and after test.

Panel Spalling.

Temperature changes during repeated heating and cooling can develop


stresses in the refractory which in turn can cause cracking. In this test, the
ability of a refractory to withstand repeated temperature changes is deter-
mined. The brick is heated to a given temperature on one side in a wall
and then artificially cooled with a fan. Each brick is weighed before and
after testing and the result is expressed in percent loss in weight. The
basic procedure, a somewhat complicated test requiring special furnaces and
control apparatus, is covered in ASTM C38. Briefly, it consists of
heating a panel of bricks, at least 18 in.5 then rapidly cooling the panel in
a blast of air and water spray for a specified number of cycles.
THE REFRACTORY

Thermal Expansion.
In contrast to the reheat test in which permanent volume changes of a re-
fractory are determined, the so-called reversible expansion of a refractory is
indicated by this test. In other words the results, expressed in percent
change in length, indicate to what extent a refractory expands when heated
to any given temperature. Subsequent cooling of the refractory will cause
it to return to its original dimensions.
ASTM vs. DIN.
All the discussions so far have been concentrated on testing methods in
accordance with ASTM procedures, i.e., procedures as they apply to the
North American refractory industry. But, foreign refractories are being
used also in the United States, and their specifications and data are, in
many instances, quite different. For this reason, a short discussion of the
differences between the European and American methods for the more
important tests is necessary.
European (DIN) testing is done on smaller cylindrical sample specimens
(50 x 50 mm) whereas American testing is done on whole bricks. This
clearly establishes an advantage for the American method because it allows
observation of the behavior of an entire brick. But, at the same time the
American method is more time consuming. The following list compares
the two methods for specific tests.
Chemical Analysis: No difference between DIN and ASTM.
Apparent Porosity: Values given by DIN are usually slightly higher
than ASTM results obtained.
Cold-Crushing Strength: No difference between DIN and ASTM
results.
Pyrometric Cone Equivalent: No significant difference between the
Seger (European) and the Orton (ASTM) test cones results.
Modulus of Rupture: DIN results tend to be higher than ASTM
results.
Refractoriness Under Load (Hot Load Test): No correlation exists be-
tween these two methods because the testing methods are different
from each other and, hence, no comparison can be made. DIN
tests for both ta, the starting temperature at which the sample is
being deformed by more than 0.3 mm and for tee, the temperature
at which the sample is being deformed by more than 10 mm.
THE ROTARY CEMENT KILN

Reheat Test: Each method uses different temperatures at which the


sample is being tested.
Panel SpuZZing: No comparison can be made between these test
results because the methods are different from each other.

3.4 TYPES OF REFRACTORIES

Refractory replacements and problems occur mainly in the burning and


the cooling zones of a rotary kiln, where the highest temperatures exist.
Linings in the calcining and heating zones very seldom have to be replaced
and usually do not represent a problem. For this reason, the following
discussion is concerned with refractory for lining in the burning zone.

TABLE 3.2
RELATIONSHIP BETWEEN PROPERTIES OF A
REFRACTORY AND ITS RESISTANCE TO ATTACK

Property to ASTM
be improved test method Required change in test result
Refractoriness c24 Higher melting point.
Higher PCE value.
Decreased percentage of impurities.
Decreased percent deformation.
stntctnral strellgth Cl6 Higher deformation temperature.
under load at high
k-
Volume stability at Cl13 Decreased linear change.
high temperaWe
Spalling resistance c38 Decreased panel spalling.
Higher thermal conductivity.
Lower modulus of elasticity.
Lower tkrmal expansion.
Higher cold compressive strength.
Slag resistance Higher density.
Abrasion resistance Higher cold compressive strength.
Higher density.

To indicate an improvement in the refractory property listed in Column 1, the test


result shcuid change as shown in Column 3.
THE REFRACTORY

The relationship between certain desirable characteristics of a refractory


and properties indicated by tests is indicated in Table 3.2. This table can
serve to assist in selecting a refractory. However, manufacuring methods,
size and shape of the units, and the type of refractory all have an important
share in influencing its ability to perform satisfactorily. For example, a
high-alumina brick can be manufactured by either the dry press or the stiff
mud method, and a brick made by the dry press process will have distinctly
different properties from a brick of identical composition manufactured by
the stiff mud process. Manufacturing methods affect the properties and
behavior of basic refractories of equal compositions also.
For the purpose of this discussion, refractories commonly used in the
burning zones can be classified into two main groups: the alumina-silica
group and the basic group.
The Alumina-Silica Group.
As the title indicates, the two main components in these refractories are
alumina and silica. With some limitations, an increase in the alumina
content (which correspondingly reduces the silica content) will result in
higher mfractories.*
In addition to the improvement in refractoriness resulting from an in-
crease in the alumina content, spalling resistance is improved (content of
fluxing materials is decreased), conductivity and strength am enhanced and
the refractory is more resistant to chemical attack. However, an important
feature is that the high-alumina brick have a greater reversible expansion, a
factor that requires special consideration during installation of the brick.
Also, the slag resistance is lower than that of basic brick.
As stated previously, manufacturing processes affect refractory charac-
teristics. Because of this a dry press high-alumina brick is preferred in
applications where uniformity of size and good spalling resistance are im-
portant, but a stiff mud brick would be recommended when high strength
and abrasion resistance are sought,

The Basic Group.


Basic refractories are manufactured mainly from per&se (a dense
crystalline magnesia), dead burned magnesite, and chrome ore. For rotary
kilns, the majority of basic bricks used fall into the magnesite-chrome
~rac&xiness: In refractories the capability of maintaining a desired degree of
chemical and physical identity at high temperatures and in the environment and con-
ditions of use (ASTM (271).
THE ROTARY CEMENT KILN

classification in which periclase makes up the largest portion of the


composition, in contrast to chrome-magnesite brick with chrome ore as
the dominant material. Basic refractories have a greater resistance to
chemical attack from ashes, slag, etc. at high temperatures, but have a
poorer spalling resistance compared to alumina bricks. Generally
speaking, the periclase in a basic refractory is responsible for high
refractoriness and volume stability, and chrome supplies spalling resistance
and hot strength to the brick. Although not conclusive, the percentages of
per&se and chrome give an early indication of spalling resistance and
other properties of a basic brick. Basic refractories are preferred in the
burning zone of a rotary cement kiln because they take on a coating more
rapidly and hold the coating better than an alumina refractory. A coating
properly formed over basic brick exhibits adherence characteristics quite
different from those of a coating formed over a high-alumina refractory, as
the coating will fuse with the surface layer of the basic brick but will only
adhere without a strong bond to high-alumina brick. Because the coating
is fused to the basic brick surface, there is a slight hazard to be considered,
for if the coating is lost, part of the brick may be lost along with it.
The third type of refractory used in the burning zone is Dolomite
bricks. These bricks are mainly composed of CaO and MgO and have a
very close chemical affinity to the kiln feed. The big advantage of these
liners is that they form a coating very rapidly once the kiln is brought to
operating temperature. Because of their high heat-transfer coefficient, these
bricks will show very high shell temperatures before such a coating is
established and it is not uncommon to observe a faint dark red shell color
in the early stages of a kiln start. The ideal location for placement of these
types of liners is in the center of the burning zone and away from the kiln
till%.
The CaO component in this brick, however, has a tendency to hydrate
when exposed to moisture in the air. Cam must be taken to make sure
these bricks do not come into direct contact with moisture in storage and
protective measures are necessary to shield the bricks from humidity during
long kiln shutdowns. Some Dolomite brick manufacturers impregnate the
bricks with tar to protect them during shipment and storage. Another
manufacturer vacuum packs their product in airtight aluminum foil to
achieve the same result. The thing to remember is that once a pallet has
been broken open, the likelihood exists that these bricks will disintegrate
in a short period of time.
When Dolomite linings are left in the kiln for a long period of time
during a winter shutdown, proper precautions must be taken so that the
THE REFRACl-ORY

lining remains unharmed. First, the coating should not be removed during
this downtime. Second, the kiln feed present in the kiln at the time of the
shutdown should also be left in and used to cover the coating/lining in the
burning zone. This partly calcined feed acts as a dessicator to absorb any
humidity present in the kiln. Third, the lining (coating) should be sprayed
with diesel oil at weekly intervals to prevent penetration of moisture to the
lining. The author knows of one cement plant that regularly shuts down
the kiln for up to 5 months each winter and is successful in maintaining
the integrity of the used, Dolomite lining. But, this appears more the
exception than the rule. One might ask the question, why bother with
these types of bricks at all? The advantages to their use are: the cheaper
price of Dolomite bricks compared to their magnesite-chrome counterparts
and the aforementioned rapid formation of new coating which is desirable
in kilns that burn a tough, difficult-burning mix.
The newest type of burning zone liner for cement kilns is the Spinell-
bonded brick. Recently introduced in the Japanese cement industry, this
liner has shown some remarkable improvements in service life for it is
reported to be as high as 1.5 to 2 times the life of high MgO-Cr liners.
Spine&bonded bricks are being offered now by almost every brick manu-
facturer and they all show chemical composition of around 10-U%
alumina and 80-85% MgO. But these bricks are still in the development
stages and each manufacturer’s Spinell-bonded brick has its own
characteristics as is the case with all other product lines. Prices for these
bricks are correspondingly about 5&100% higher than conventional basic
liners. Because of these high prices, Spine&bonded bricks are normally
installed in such places in the burning zone where everything else has
failed.

3.5 REFRACTORY SHAPES

On the North American continent, rotary-kiln blocks, arches, and


wedges am the most common refractory shapes used to line rotary kilns.
In countries using the metric system of units, VDZ and IS0 shapes are
used. The following data will familiarize the reader with the dimensional
differences between these shapes. It is important to note that dimension
“a ” , i.e., the back cord, is the face of the refractory that is in contact with
the kiln shell. All shapes are installed so that the given dimension “P
forms a parallel line to the kiln axis. Dimension “h” indicates the lining
thickness.
THE ROTARY CEMENT KILN

hfp h@ hQ
Rotary kiln
block (RKB) Arches Wedges

VDZ shapes IS0 shapes

TABLE 3.3
DIMENSIONS IN THE ENGLISH SYSTEM OF UNITS
(dimensions given in inches)

RKB Arches wedges V D Z IS0 x13


,
9 9 12 4 3.5 3 4 3.5 3 var. 4
f v v v v v V v v v VX.
C 2.i
1 ; ; 4’ 9’ ii ii : i ii 7.8 7.8
h 6 9 9 66 6 9 9 9 VU. VBT.

Note: VDZ and IS0 shape-s are usually manufactured for 6 3/10 in., 7 in., 7 7/8 in., 8
718 in., and 9 4/5 in. lining thickness (h).
THE REFRACTORY

TABLE 3.4
DIMENSIONS IN THE METRIC SYSTEM OF UNITS
(dimensions given in millimeters)

RKB Arches Wedges I?02 IS0 n I3

a 229 229 305 102 89 76 102 89 76 ViU. 103


b v v v v v v v v v V
7;:2
f 1B 42 Iti2 49 29 2;9 1:2 Ii2 I:2 198 198
h 152 229 2 2 9 152 152 152 229 229 229 var. var.

Note: VDZ and IS0 shapes are usually manufactured for 160, 180, 200, 225, and 250
mm thick linings (h).

3.6 NUMBER OF BRICKS REQUIRED


PER RING
Outside the United States, it is customary to use two different shapes of
bricks with different backcords (a) to complete a full circle of the kiln cir-
cumference. Experience has shown that this produces a superior fit of the
refractory to the shell particularly when the kiln shell is slightly out of
round.

a) for RKB, arches and wedges


for basic lining for aluminum brick

12 (14t 12dlc
?I = n =
a + 0.059 a + 0.039

n = number of bricks per ring


a = back cord (in.)
d = kiln diameter (ft)

b) for VDZ shapes

Consult manufacturer for number of bricks per ring of each shape


It!Cpired.
THE ROTARY CEMENT KILN

c) For I.S.O. shapes (7rf3)

I.S.O. shapes have a uniform backcord of 103 mm. With an expansion


insert of 1 mm, the cord length becomes 104 mm which is equivalent to
ld3, explaining the reason for identifying these shapes by this nomencla-
ture. With IC a constant in the brick backcord and the circumference of the
kiln shell, the calculation for the number of bricks required per circle be-
comes simple:

n = 10000(0.0333) = 33.330

where
D = internal kiln shell diameter (m)

Example: How many bricks are required in I.S.O. shapes to complete a


circle on a 4.8 m diameter kiln?
Answer: (33.33) (4.8) = 160 pieces.

3.7 NUMBER OF BRICKS REQUIRED


PER UNIT KILN LENGTH
a) when dimension “I” is expressed in millimeters and kiln length in
meters:

I N=
Brick
m
RKB
9.6n
Arches
4.3n
Wedges
6.5n
VDZ
5n
IS0
5n I

b) when dimension “I” is expressed in inches and kiln length in feet:

Brick RKB Arches Wedges VDZ IS0


N=
ft 3n 1.33n 2n 1.524n 1.52.h

n = number of bricks per ring


THE REFRACTORY

TABLE 3.5
KILN DIAMETER CONVERSION TABLE

ft mm ft mm ft mm

8.84 2900 12.00 3937 15.54 5100


9.00 2953 12.19 4000 15.85 5200
9.14 3000 12.50 4101 16.00 5250
9.45 3100 12.80 4200 16.15 5300
9.50 3117 13.00 4265 16.46 5400
9.75 3200 13.11 4300 16.50 5414
10.00 3281 13.41 4400 16.75 5500
10.06 3300 13.50 4429 17.00 5577
10.36 3400 13.72 4500 17.07 5600
10.50 3445 14.00 4593 17.37 5700
10.66 3500 14.02 4600 17.50 5742
10.97 3600 14.32 4700 17.68 5800
11.00 3609 14.50 4757 18.00 5900
11.25 3691 14.63 4800 18.29
11.27 3700 14.93 4900 18.50 6069
11.50 3773 15.00 4922 18.59 6100
11.58 3800 15.24 5000 18.90 6200
11.89 3900 15.50 5085 19.00 6234

3.8 REFRACTORY CONSUMPTION REPORTING

Finally, mention must be made of methods of reporting refractory con-


sumption. It is easy to fall into the trap of reporting refractory service life
in terms of days, months, or years. Unless the kiln has operated con-
tinuously during this period of so-called “days service life” there is not
much use in comparing such figures with other periods or kilns. A kiln
manager should, if possible, find the time to express refractory consump-
tion in terms of kilograms refractory per ton of clinker (or Ib/sh.t.Cl.)
produced. More paperwork is required but the data thus obtained is much
more meaningful to the study of individual refractory types. Likewise,
proper record keeping on brick charts is an absolute requirement for any
plant that is seriously concerned about improving refractory service life.
THE ROTARY CEMENT KILN

References

1. Z. Clausen, C.F. Jan., 1%2 ‘The Evolutior~ of the Cement Kiln, a


Historical !SketcV Journal, PCA Research and Devebpment
LUbOfOtO&?S.
2. Budnikov, P.P. ‘The Technology of Ceramics and Refractories”
M.I.T. Press, Mass. Inst. of Technology.
4.

Fuels
Fuels used in burning a rotary kiln are classified into three groups:
solid, liquid, and gaseous fuels. In many kilns a combination of two
different fuels are used.

4.1 SOLID FUELS

Solid fuels used in rotary kilns are generally of the following types and
properties:

M0is1we Volalilc Fixed Ash Heating valw


Group molier co&on
m @Jo) (W (%6) W&i Btib

Anthracite 23 3.1 87.7 6.9 31308 13540


Semianthracite 2.2 12.4 67.4 18.0 28,435 12,270
Bituminous low-v01 3.5 18.2 74.4 3.9 33,858 14350
Bituminous med-vol 3.1 25.0 66.8 5.1 3333 14290
Bituminous high-vol A 2.6 30.0 58.36 9.1 31,670 13510
Bituminous high-vol B 8.2 36.1 48.7 7.0 2429296 12,160
Bituminous high-vol C 12.1 40.2 39.1 8.6 26$714 11,480
Subbituminous A 16.5 34.2 38.1 11.2 22,665 9,740
Subbituminous B 23.2 33.3 39.7 3.8 21,920 9,420
Subbituminous C 24.6 27.7 39.9 7.8 20,035 8,610
Lignite 34.8 28.2 30.8 6.2 16,778 7210
Delayed coke 13.0 1.2 33,974 14,600
Fluid coke 5.0 2.0 34,440 14,800
THE ROTARY CEMENT KILN

Subbituminous coal, lignite, oil shale, and other low-grade, solid fuels
are fired predominantly in the flash calciner since these low heating-value
fuels are unsuitable to sustain good combustion conditions in the burning
zone. As a matter of fact, one of the most important advantages of a pre-
calciner kiln is that low-grade fuels can be used for firing in the auxiliary
furnace of the precalciner. Petroleum coke is used either at the precalciner
furnace or intermixed with other fuels in the burning zone. Several pre-
heater kilns in various parts of the world are being used to dispose of old
automobile tires, others use wood chips, and even garbage for introduction
to the back end of the kiln. This method serves two purposes simul-
taneously. It does away with an environmental problem of waste material
disposal and at the same time delivers valuable heat to the kiln which
otherwise would have to be suppplied by the more expensive conventional
fuels. Petroleum coke possesses a much higher ignition temperature
[=590 C (1100 F)] and is therefore suited for intermixing with raw feed in
a suspension preheater kiln. In effect, this method upgrades a preheater to
precalciner status without the need for installation of an auxiliary firing
system at the preheater tower. There is one drawback in using petroleum
coke in any rotary kiln, namely its unusually high sulfur content.
One of the inherent disadvantages of coal is not only its explosive
potential when finely ground but also its tendency to ignite spontaneously
while in storage. High-volatile, high-sulfur, high-moisture coals are es-
pecially vulnerable to such ignition. Special precautions have to be taken
when stockpiling coal to prevent undesirable temperature rises in the pile.
Common precautions taken as standard procedures arc:

a) Building a coal pile in layers of 2-3 ft and compacting these


layers with a bulldozer or roller before the next layer is applied.
b) No foreign materials such as rags, pieces of wood, or paper
should be present and the storage area should contain a concrete
floor with proper drainage.
C) Testing wells (closed-ended steel pipes inserted into the pile)
should be strategically spaced over the entire pile to allow for
temperature monitoring by means of thermocouples or
thermometers.
d) For long-term, strategic, coal reserve piles, coal should be
screened with the maximum amount of fines removed, i.e., store
only coarse lump coal if possible.
e) When a smothering “hot spot” is detected, do not apply water in
FWFLS

an attempt to put the “fire” out. Instead, dig it out and repack the
pile.

Coal firing on rotary kilns is done by either a direct-, semidirect-, or an


indirect-fired system In the direct-fire system, the coal is dried, ground,
and immediately conveyed into the burning zone for firing (see Fig. 2.1).
Grinding is being done in an air-swept bowl or roller mill with the particle
size classification being done within the mill itself. Hot, excess cooler
air, tempered down to safe levels is used for drying and conveying of the
coal. This system is normally quite safe, easy to control but has the draw-
back that a large amount of relatively cold primary air must be used for
firing. This makes this system less efficient than an indirect-fued one. It
also limits the capability of the operator to change the flame shape at will.
Most wet- and dry-process kilns, i.e., the older types, are being fried by
this method.
Semidirect-fired systems mostly use an air-swept ball mill from where
the pulverized coal is conveyed to a small cyclone (holding bin) above the
primary air pipe. The coal then discharges through a variable-speed rotary
feeder directly into the primary air pipe from where it is being blown into
the kiln (see Fig. 2.2). Since these holding bins are usually relatively
small, the mill, just as in the case of the direct-fling system, must be op-
erating at all times while the kiln is running. When compared to the
direct system, semidirect firing requires less primary air to be us& can
operate at higher primary air temperatures, and can still supply the kiln
with coal when the coal mill is shut down for up to 10 min.
In the indirect-fired system the coal is dried and ground in a roller or ball
mill system that is separate from the actual kiln firing system The
ground coal is stored in a large silo from where it is withdrawn at metered
rates to the primary air pipe (see Fig. 2.3). This grinding system does not
have to be operated continuously as the capacity of the coal silo is usually
sufficiently large to hold a supply of coal for up to 10 h. This is the type
of system that most new cement kilns employ. Several kilns can be fired
from this single grinding plant and coal rates, to the kiln, can be adjusted
instantaneously by the operator. It also allows for optimum primary air-
flow control. This indirect system is the most complex of the three
systems. It needs special controls, safety devices, and a dust (coal dust)
collector because of the presence of vast quantities of finely ground coal.
It is important to strongly emphasize to kiln operators how important
it is to keep close control over any coal-firing system These systems
THE ROTARY CEMENT KILN

have been designed for safety, am safe, and can remain so provided they
receive the proper attention and maintenance. There are no shortcuts that
can be used here: standard operating procedures must be fully understood
and adhered to for the simple reason that fine pulverized coal can be an
ingredient for a highly explosive mixture when certain conditions prevail
There is no reason to have any fear of these systems as long as the opera-
tor fully understands what must be controlled and knows exactly what to
do when something goes wrong. An operator must show respect toward
coal ftig and should periodically review the operating procedures to have
himself prepared for any eventuality. Most important of all, if an operator
has some doubts about the accuracy of any instrument reading in the coal
grinding and firing system, or when he knows something is not function-
ing properly, he must have it fixed right away.

4.2 LIQUID FUELS

Liquid fuels used for combustion in a rotary kiln are almost exclusively
of the Bunker B or Bunker C class. These are residual oils from refineries
after the more volatile products in the oil have been removed Because
both types fall into the class of heavy oils, they have to be preheated
before they can be pumped and atomized. Steam plants usually deliver the
necessary heat to the heat exchangers where the fuel is raised to the
temperature that ensures the correct viscosity.
To obtain good combustion, fuel oil has to be atomized, which means
the oil has to be broken into small droplets to promote easy combination
with the oxygen. This is done by means of an atomizer nozzle in which
the oil is forced through an orifice at high pressure, creating a turbulent
motion. The orifice size determines the pressure of the fuel at the bumer-
tip, which in turn influences the flame shape. A small orifice results in
higher fuel pressure and less fuel passing through the burner than a large
orifice. Because certain high and low limits are set for a given orifice, it is
often necessary to exchange the orifice for one with a different size of
opening when a large change in fuel rate is required.
Fuel pressure is of utmost importance for the shape of the flame and
consequently has to be frequently checked by the kiln operator. If the
pressure, for example, is too low, complete combustion of the poorly
atomized oil particles cannot take place, resulting in impingement of the
unburned fuel on the coating and the feed bed.
4.3 GASEOUS FUELS

Although other gaseous fuels such as liquefied petroleum gas, hydrogen


and coal gas are used in some industries, gaseous fuel fling of rotary kilns
in the cement industry is done almost exclusively with natural gas, one
reason being that it is the cheapest of all gaseous fuels. Natural gas offers
several advantages for the kiln operator, as it needs no preparation such as
drying, grinding, or preheating, and combustion takes place readily once
the gas has been mixed with the proper amount of air and ignition tem-
perature has been reached. The burning zone atmosphere is much “cleaner’
because a natural gas flame is not as luminous as an oil or coal flame. A
natural gas-fired burning zone appears “hotter” to the operator than one
with other fuels, making a short adjustment period for the kiln operator
necessary whenever firing is switched from another type of fuel to natural
Em.
Another important advantage of natural gas firing is the need for very
little or no primary air, thus more valuable hot secondary air can be used
for combustion in the kiln.

4.4 COMPOSITION AND HEAT VALUE

A basic knowledge of heat generation and fuel efficiency is essential for


any kiln operator, who should remember that the cost of fuel is one of the
major expenditures in the manufacturing process of cement. Efficient use
of the fuel depends in a large measure on the kiln operator’s skill.
Composition and heat value of fuel are of significant importance to
everyone concerned with operation of a rotary kiln, as they are the basis for
heat transfer calculations and for investigation of the fuel efficiency of a
kiln. Fuels differ considerably in the amount of heat they give out when
burned, hence it is desirable to measure their heating values by means of
an instrument called a calorimeter, expressing the heating value as the
number of calories or British thermal units (Btu) given off when a mea-
sured amount of fuel is burned. One gram-calorie, or small calorie, is the
amount of heat required to raise the temperature of one gram of water one
degree centigrade. Similarly, a kilogram-calorie, or greater calorie, is the
amount of heat required to raise the temperature of one kilogram of water
one degree centigrade. In the English units, one Btu is the amount of heat
THE ROTARY CEMENT KILN

required to raise the temperature of one pound of water one degree Fahren-
heit. In North America, heating value of fuels is usually expressed as the
number of Btu per pound for coal, Btu per pound or gallon for liquid fuels,
and Btu per cubic foot for gaseous fuels. See Table4.1. Metric units may
be either gram-calories or kilogram-calories per kilogram or liter. Con-
version factors are given in the appendix.

TABLE 4.1
COMPOSITION AND HEATING VALUE OF FUELS

SOLID FUEL
Pulverized coal
Volatile
Carbon Hydrogen Oxygen Nitrogen Sulfur Ash matter Btu/lb
78% 5% 6% 1.5% 1% 8.5% 30% 13,000

LIQUID FUEL
Bunker oil C
Carbon Hydrogen Oxygen Nitrogen lb/gal BNgal Btu/lb
86% 12% 1% 1% 8.1 146,500 18,700

GASEOUS FUEL
Natural gas
Other
BltU&t3
Mit?e
0 EsY0 Ni%!?m hydTPOns
0

With the recent world-wide (except USA) standardization of the system


of units, fuel heating values are now expressed in kiloJoules per kilogram
@J/kg) or per cubic decimeter (kJ/dm3). Conversion factors are given in
the appendix.
Divergent views still exist about the manner in which specific fuel con-
sumptions are expressed. If it is stated that a particular kiln consumes
3722 Id/kg of clinker (3.2 million Btu/sh.ton), then more information
would be necessary. Determination of whether the gross or net heating
value of the fuel was used and whether the fuel consumption was under
steady-state kiln operating conditions or based on month-end fuel inventory
figures would have to be made. If the cement industry would adopt the
“NET HEATING VALUE/STEADY STATE” method, then efficient and
accurate determinations of fuel consumption could be made. Gross heating
FUELS

values are the heat produced by a unit quantity of coal in a bomb calori-
meter. The net heating values are based on both the heat produced at con-
stant atmospheric pressure and the water in the combustion product
remaining in vapor form. Net heating values can be calculated from the
results of the calorimeter test by subtracting the hydrogen component as
follows:

a) for Metric units

=4lti = o+ross - 5150 HJ 4.187 W&s)

b) for English units

NET value = GROSS value - 99270 Hz (Btuflb)

The gross heating value is approximately 3.5% higher than the net
heating value. It is quite possible that a particular kiln might show a spe-
cific fuel consumption of 4886 kJ/kg (4.2 MBtu/sh.t.) in steady-state
condition but the month-end figure reported might show this as 5375
kI/kg (4.62 MBtu./sh.t.) because these month-end figures include the clink-
er and fuel losses due to frequent kiln stops and starts or material losses
that might not be associated at all with the operation of the kiln itself. It
is easy to get on the wrong track in the matter of fuel conservation when
coal and clinker storage areas are inadequate to prevent these materials from
being blown by the wind over the countryside, or washed down the road
during heavy downpours. These areas can not be overlooked in a fuel
conservation program. It is therefore essential to check the specific fuel
consumption under steady-state conditions and then compare this with the
specific fuel consumption that is derived from the month-end inventory
figure. If the latter is higher than 3% from the former, it is necessary to
take a serious look at the existing clinker and fuel handling and storage
systems. Attention to this can save the company many thousands of
dollars every year.
5.

Combustion

Before entering into a detailed discussion on the combustion conditions


in the kiln, it is appropriate to review some fundamental laws related to
gases and combustion. Knowledge of these laws are essential in order for
any kiln operator to understand the duties involved in his job. When mal-
functions and unusual problems occur, such as chain fires, rapid and un-
usual temperature rises in the fuel or firing system, an understanding of
these laws combined with common sense can make the difference between
a potentially serious, out-of-control condition and a safe solution.

5.1 GAS LAWS

All gases, including the hot gases in a kiln, behave in a certain manner
under external influences. A perfect gas is one that obeys very closely cer-
tain physical laws. For all practical purposes, the gases under considera-
tion in this chapter may be assumed to be perfect gases. Adjustments to
the air circuitry of a rotary kiln are part of the responsibility of the kiln
operator, hence a basic knowledge of these laws is necessary to assist him
in making the correct adjustments.
Before entering into a discussion of the gas laws, the reader is reminded
that pressures and temperatures used in the following equations are
absolute . Absolute values are determined:

a) by adding 460’ to the temperature in degrees Fahrenheit, e.g.,


600F = 520” absolute.
COMBUSTION

b) by adding atmospheric pressure to gauge pressure; while


atmospheric pressure varies with altitude and weather conditons, it
is sufficiently accurate for these computations to use 14.7, i.e.,
50 psi gauge = 64.7 psi absolute.

In a like manner, CGS units are obtained by adding 273 ;o the


temperature in degrees centigrade, and 76 to the pressure in centimeters of
mercury.

Boyle’s Law.

In 1662 Robert Boyle expressed this law as follows:


Under constant temperature, the volume of a given mass of gas varies
inversely as the pressure upon it. In other words, pressure times volume,
at constant temperature, is a constant expressed mathematically
PlVl = P$5

where:
PI= initial pressure (absolute)
P2 = final pressure (absolute)
VI = volume under PI
V2 = volume under P2

Example: The volume of a certain gas, when measured by an instrument


indicating a barometric pressure of 72 cm of mercury, is 43,000 ft3- What
will be the volume at standard presure (76 cm Hg) while the temperature
remains constant?
Substituting in Eq. (5-l)
72 x43,000 = 76xV2

v, = 43,000 ; = 40,737 ft3

Charles’ Law.

Similar to the discovery of Boyle, Jacques Charles in 1787 discovered


that the volumes occupied by a given mass of gas at different temperatures
are proportional to the absolute temperatures of the gas provided the
THE ROTARY CEMENT KILN

pressure remains constant


All gases expand when the temperature is raised. If the temperature of a
gas is raised 1 F, it will expand l/460 of its original volume. In CGS
units, a temperature rise of 1 C will result in an expansion of l/273 of the
original volume.
Mathematically

Vl v2
-=-
G-2)
=1 =2

where:
T1 = initial temperature, absolute
T2 = final temperature, absolute
VI = volume under T1

Example: The volume of a gas when measured at a temperature of


450 C is 5700 m3. What will its volume be at 520 C?
Reduce the Celsius reading to absolute (Kelvin) temperatures:

45OC=450+273 ~723 K
52OC= 520+273 =793K
5700 v,
-=-
723 793

793 = 6252 m3

Gay-Lussac’s Law.

Finally, there was Joseph Gay-Lussac who, early in the 19th Century,
delved further into the action of gases, and developed the law that bears his
name, which states that the pressure exerted by a given mass of gas will
increase in proportion to the temperature if the volume is held constant.
All units again are in absolute values.
COMBUSTION

PI p2
=-
G-3)
T1 T2

in which pressure and temperature units am as previously described.


Example: The pressure of a gas when measured at 900 F is 72 cm of
mercury. What wiIl be the pressure if the temperature is raised to 1050 F
and the volume remains constant?
Reduce the Fahrenheit readings to absolute temperatures:
900F=900+460=1360A
1050 F = 1050 + 460 + 1510 A
Substituting the values in Eq. (5-3)
72 P2
1360 = -
1510

0.05294 x 1510 = 79.9 cm Hg

A General Law.

Now reviewing the above three basic gas laws, it becomes apparent that
a general law can be stated, based on all three of the basic laws. The gen-
eral equation is:

Pl”1 P2”2
=-
(5-4)
Tl T2

in which PI, VI, and T1 are the original pressure, volume, and
temperature, and P2, “2, and T2 are the final values, with temperatures and
pressures expressed in absolute units.
Example: A gas measured 45,000 fts at 900 F under a pressure of 75 cm
Hg. What will be the volume at 1050 F at a pressure of 76 cm?
First reduce the Fahrenheit readings to absolute:
900 F = 900 + 460 = 1360 A
1050 F = 1050 + 460 = 1510 A
Substituting in Eq. (54)

75 x 45,000 76 x V,
=-
1360 1510
THE ROTARY CEMENT KILN

75 x45,ooo x 1510
v, = = 49,290 ft3
76 x 1360

These gas laws can help a new operator to understand some of the fun-
damentals in kiln control. They are the solutions to some common basic
control problems an operator might be confronted with on an almost daily
basis.

5.2 THE COMBUSTION REACTION

Successful operation of a rotary kiln requires an adequate source of


heat that will first raise the kiln to the desired operating temperature, and
will then maintain this temperature by compensating for the various heat
losses occurring in the kiln system, including the heat required for the
process. The required heat is obtained by combustion of the fuel, a
chemical rection in which carbon, C, hydrogen, Hz, and sulfur, S, in the
fuel combine with oxygen, 02, in the air. To obtain combustion, two
requirements must be fulfilled:

a) sufficient oxygen must be present to mix with the fuel, and


b) a certain temperature must be maintained to ignite the fuel-
oxygen mixture.

An operator must always remember the triangular relationship that leads


to combustion:
FUEL

If any one of the three links is missing, no combustion will take place. If
there is not enough air in the kiln, there will be no proper fire. Likewise
in a dry or wet kiln, too much heat and air in the chain section will result
COMBUSTION

in chain fue because chains contain carbon which acts as fuel. In order to
stop a fire, the elimination of one of the components of the triangle is
necessary, e.g., choking off the oxygen (air) supply. This will be dis-
cussed in greater detail later on. In addition, firing conditions inside the
kiln must be such that the fuel particles undergo complete combustion
while the fuel is still in suspension in the kiln atmosphere.

5.3 THE STANDARD COAL FACTOR,


COMBUSTION AIR REQUIREMENTS

To determine the approximate combustion air needed to bum a given


unit weight of coal, the formulas given below can be used when no ulti-
mate coal analysis is available. The combustion air requirements here
include 5 percent excess air.

a) English units b) Metric units

100-a b 100-a B
SCF = loo - SCF = (5-5)
12,600 100 7000

lb air/lb coal = 10.478 SCF kg air/kg coal = 10.478 SCF

where:
SCF = standard coal factor
= percent moisture in coal (as fired)
i = heat value of coal (Btu/lb as fired)
B = heat value of coal (k&kg as fired)

Fuel ignition temperatures, the third component in the combustion


triangle, are approximately:

“C OF
for coal 250 480
for fuel oil 200 400
for natural gas 550 1050
for petroleum coke 620 1150
THE ROTARY CEMENT KILN

Here it can be seen why gas- and coke-fired kilns tend to have their
ignition point of the flame much further back in the kiln than coal- or oil-
fired kilns. In addition, combustion in a kiln requires that sufficient time
must be available to accomplish complete combustion while the fuel is in
suspension in the kiln atmosphere.
When complete combustion takes place, carbon dioxide (COz), water
vapor (H,O), and sulfur dioxide (SO,) are formed:

c+o, +co,
2Hz + 0, + 2H20
s+q +sos

These equations for complete combustion indicate that when fuel is


properly burned, 1 part of carbon will combine with 1 part of oxygen to
form the combustion product carbon dioxide. However, when incomplete
combustion takes place, then instead of carbon dioxide the combustion
product is carbon monoxide:

2c+oz+co

Oxygen needed for combustion originates from the air which is forced
into the kiln. Air consists mainly of 78 volumes (76% by weight) of
nitrogen and 21 volumes (23% by weight) of oxygen. Thus, it is
necessary to introduce approximately 5 volumes of air for each volume of
oxygen needed to obtain complete combustion. The nitrogen contained in
the air does not enter the combustion process, only the oxygen reacts with
the carbon, hydrogen, and sulfur to form the combustion gases.

Effect of Kiln Air on Combustion Efficiency.

It should now be apparent to the readers that control of the air supply
for the kiln is as important as control of the fuel rate, because each is
dependent on the other in their effect on combustion. In the following it
will be shown that too little air (deficiency) as well as too much air
(excess) is harmful to the economical operation of a rotary kiln.
When combustion is incomplete because of a deficiency of air, approxi-
mately 4500 Btu are released when one pound of carbon is burned to
carbon monoxide. However, under conditions of complete combustion in
burning the same amount of carbon to carbon dioxide, 14,500 Bt,u will be
COMBUSTION

released, a difference of 10,000 Btu between complete and incomplete com-


bustion, a clear indication of the importance of having at all times enough
air available to accomplish complete combustion of the fuel. It should be
pointed out that a rotary kiln, for safety reasons, is never allowed to be
operated with such a deficiency in air supply. There is no room for any
compromise in such a situation; the operator must do everything in his
power to again obtain complete combustion of the fuel within seconds
after the first indication of incomplete combustion is given. It is defi-
nitely the wrong procedure to increase the fuel rate at a time when a de-
ficiency of air is already present and the burning zone starts to lose
temperature.
More common is a rotary kiln operating with a large percentage of ex-
cess air, usually resulting from inefficient attendance of the kiln by the
operator. The question might be raised now why too much air is also
harmful to the efficient operation of a kiln, the reasoning being that with
plenty of air available a total burning of the fuel will with certainty be
secured. However, operating a kiln with a large percentage of excess air
increases heat losses to the rear of the kiln and lowers the flame tem-
perature. Furthermore, any excess air not required for combustion con-
sumes valuable heat because this air too will be raised to the operating
temperature of the kiln. In short, by consuming heat to raise the tem-
perature of the excess air, less heat will be available for the actual burning
process of the material in the kiln.
G. Martin,3 in an extensive study of the thermodynamics of rotary
kilns, has calculated that for every 1% of free oxygen present in the kiln
exit gases, there is a loss of 0.4449 tons of fuel for every 100 tons of coal
burned. That is, 0.4% of the heat introduced into the kiln by the fuel will
be lost for each 1% free oxygen in the kiln exit gases. Consider, for
example, a kiln producing 2040 tons clinker per day at an average specific
consumption of 4.4 MBtu/ton clinker when the exit gas oxygen content is
1.5%. Burning the same amount of clinker per day but with 4.5% oxygen
in the exit gas would result in a loss of 120 MBtu because of the excess
air admitted to the kiln. At a common price of $1.80 per MBtu, roughly
$67,000 per year would be lost.

Determining Excess or DeJiciency of Air.

The foregoing discussion naturally leads to the question: How does an


operator know when the kiln is operating with too much or too little air?
THE ROTARY CEMENT KILN

To assist the operator, analyzers are available that continuously deter-


mine the amount of oxygen and combustibles (carbon monoxide) in the
kiln exit gases. Other analyzers, less frequently used, are available for
determining carbon dioxide content. The Orsatt apparatus is used for
making periodical analyses of gas samples to determine oxygen, carbon
monoxide, and carbon dioxide content.
The percent oxygen contained in the kiln exit gases gives the best in-
dication of the combustion condition in the kiln because this oxygen is
directly related to the amount of air introduced and the amount of oxygen
taken up during the process of combustion. We know that air contains
21% oxygen by volume. If there were no combustion reaction in the kiln,
the same percentage of oxygen would be in the air leaving the kiln. How-
ever, because combustion reactions do take place in the kiln, most of the
oxygen reacts with carbon, hydrogen, and sulfur to form the combustion
products CO, CO,, H20 and S02. Thus, when no free oxygen can be
found in the kiln exit gases no excess air has been introduced into the kiln.
The percent excess air can be calculated with the formula given by
Perry, Chilton, and Kirkpatrick,4 provided that no combustibles (CO) are
present in the kiln exit gases:

percent excess air = 21 _ o, K (5-6)

in which:
K= 0.96 for bituminous coal
0.95 for oil
0.90 for natural gas, and
02 = oxygen content of exit gases expressed as a percentage.
Fig. 5.1 shows this relationship for a typical rotary kiln.
The percent excess air can also be calculated from the results of a gas
analysis obtained with the Orsatt apparatus. The following formula is es-
pecially helpful when a gas has a content of less than 1% oxygen because
such gases usually contain traces of carbon monoxide also.

189(202 - CO)
percent excess air = O-7)
N2 - 1.89(202 - CO)
COMBUSITON

% EXCESS AIR

Fig. 5.1 Ideal operating conditions in the kiln occur when the kiln exit gas
contains between 0.7 sad 3.5% oxygen. Zone A indicates an excess of air,
resulting in excessive heat loss; Zone B indicates a deficiency of air, resulting
in the formation of carbon monoxide.

in which
02 = percentage of oxygen
C O = percentage of carbon monoxide, and
N2 = percentage of nitrogen

Incomplete combustion is easily recognized when any percentage,


regardless of how small, is indicated on the combustible (CO) recorder. As
stated earlier, a rotary kiln is never allowed to operate under conditions of
incomplete combustion which means that at no time should there be any
carbon monoxide present in the kiln exit gases.
Now consider the question of just what constitutes the ideal operating
condition for efficient combustion and kiln control. At this point, theory
and practice are apt to take divergent paths. Many viewpoints and opin-
ions, some good, some in error, have been advanced, so it is well at this
point to review the theory of perfect combustion. First, most efficient
combustion takes place when there is neither carbon monoxide nor excess
air present in the kiln exit gases: that is, the oxygen and the combustibles
THE ROTARY CEMENT KILN

recorders should both have a reading of zero. Second, with any increase of
either carbon monoxide or excess air, valuable heat is lost.
Application of this theory is of value only when perfect combustion
conditions prevail within the kiln burning zone. In practice, however, this
condition is rarely attained because many factors associated with design of
the kiln work against such ideal conditions.
A typical example that spotlights this fact is the common observation
that a small percentage of carbon monoxide can be found in the exit gases
while at the same time there is also a small percentage of free oxygen pres-
ent. This is in direct contradiction to the combustion theory that carbon
monoxide shows only after all the excess air has been used up; that is,
after the oxygen level has fallen to zero. In many instances the Orsatt
analysis or the recorders have shown combustibles at oxygen readings in
the range of 0.7%. This indicates that inefficient conditions are present
when oxygen drops below 0.7%. The operator can observe the effects of
changes in oxygen on the flame itself. If, for example, the kiln is operat-
ing at an oxygen content of 0.7% and the fuel rate is increased without an
increase in the air flow in the kiln, such an action will cause a change in
the color of the flame, the flame taking on a darker color at its outer rim, a
sure sign that the flame temperature is dropping.
Through experience it has been found that a rotary kiln operates best
when the kiln exit gases have an oxygen content of not less than 0.7% and
not mom than 3.5% under stable operating conditions. The optimum
target point is between 1.0 and 1.5% oxygen. In addition, under no
circumstances should there be any carbon monoxide present in the kiln exit
gases. The given targets and ranges for oxygen levels do not apply at
times when the kiln is in an upset condition.
New and more advanced types of gas analyzers have recently been
installed in many kilns. These advanced technology analyzers are capable
of detecting and recording minute traces of gas components such as 02,
CO2 CO, as well as SO2 and NO,. Recordings on these units are usually
in terms of ppm (parts per million) instead of percentages.

Pm %
10 0.001
100 0.01
Loo0 0.1
10,000 1.0
COMBUSTION

The thing to remember is that conventional gas recordings will only


indicate the presence of CO when it exceeds approx. 500 ppm whereas the
newer analyzers are capable of recording contents as low as 10 ppm. It is
therefore necessary to reword the above-mentioned operating rule and state
that immediate counteracting moves must be made whenever the kiln gases
contain more than 100 ppm (0.01%) CO.
The fact that both too little and too much air for combustion is detri-
mental for fuel economy, has led to the practice in some plants of main-
taining a constant fuel-to-air ratio at all times when the kiln is operating,
even reaching the stage on some kilns in which controllers automatically
adjust the total mass of air entering the kiln whenever a change in fuel rate
is carried out. Although such an approach is fully justified from the
viewpoint of combustion efficiency, it is not conducive to overall
operating stability of the kiln.
Furnaces with less complex reactions than a rotary kiln can operate
efficiently under the above principle. In the rotary kiln, however, there are
many more factors that have a direct influence on the stability of the opera-
tion. For example, an important factor that is neglected when the prin-
ciple of a constant fuel-to-air ratio is used to control a kiln, is the heat
profile through the entire length of the kiln. A rigid application of this
burning technique causes the temperature profile to change toward the back
end of the kiln.
Table 5.1 shows the results of a burning technique in which the temper-
ature profile has not been considered Examples A and C follow the rec-
ommended practice of controlling the oxygen level within a range of 0.7-
3.5%, and at the same time giving full consideration to the back-end tem-
perature. Example A shows how the fuel rate and the air rate into the kiln
are proportionally changed with succeeding increases in the kiln feed rate.
With each feed rate increase there must be a corresponding increase in fuel
rate and air flow rate as well as a rise in back-end temperature to take care
of the increased production. This, then, is the procedure for stable kiln
operation. Failure to make all these required increases could lead the kiln
into an upset. There is little doubt that such an approach as shown in
Example A will meet the approval of experienced operators.
Examples of upset kiln conditions are shown in B and C in which the
feed rate remains constant. Example B shows the technique of maintaining
a constant fuel-to-air ratio; that is, the oxygen percentage remains con-
stant This results in uniform combustion but very serious fluctuations in
burning zone and back-end temperatures. Changes in back-end temperature
THE ROTARY CEMENT KILN

TABLE 5.1
TECHNIQUES OF COMBUSTION AND KILN CONTROL

Example A: propOrtional increases in fuel rate, back-end temperature and


I.D. fan speed (+r rate) for increases in the feed rate.

Feed Burning z o n e Fuel ID. fM Back-end


rote temperature rate speed Oxygen temperature

E %z 200,000 180,000 380 395 2 1440 1480


108 2800 220,000 410 2 1520
120 2aoo 240,000 425 1560
Example B: Fuel-to-air ratio constant (oxygen constant); feed rate
constant, but kiln is in an upset.

Feed Burning zone Fuel I.Ll. fan Back-end


rote temperature rate speed Oxygen temperature

240,000 425 1560


200,oo 395 ;: 1480
96 2950 180,000 380 2:5 1440
Example C: I.D. fan speed adjusted to compensate for temperature changes
in the back end due to fuel rate changes. As in example B the feed rate is
constant but the kiln is in an upset condition.

Feed Burning zone Fuel ID. fan Back-end


rote temperature rote speed Oxygen temperature

E ~~ 240,000 200,000 385 395 i-ii 1480 1480


96 2950 180,000 405 3:5 1480

result in uneven feed preparation (drying, calcination) in turn creating


cycling operating conditons. Because the feed rate is constant, heat distri-
bution should be unchanged to maintain the same drying and calcining
conditons.
Example C is the recommended procedure for kiln upsets in which an
attempt is made to maintain back-end temperature constant. Small adjust-
ment are made to the induced draft fan speed to compensate for the temper-
COMBUSl-ION

ature drops or rises that occur in the back end due to changes in the fuel
rate.
The preceding discussion of combustion has centered on and is
applicable to firing systems where the air supply to the furnace can be in-
dependently controlled. In other words, it applies to all kilns (wet, dry,
semidry, preheater) that have one single-firing system in the lower part of
the kiln (burning zone). It also applies to precalciner kilns that have ter-
tiary air coming from the cooler for the flash furnace. The discussion on
the proper air supply to the kiln and the optimum percent of excess air
does not apply to precalciner kilns where all the air for the precalcining
chamber (flash furnace) originates from and goes through the rotary kiln
(i.e., no tertiary air duct present). In such kilns one single air-supply route
must serve two combustion processes namely the flames in the burning
zone and the flash calciner. Obviously, the excess air at the rotary kiln
back end must be much higher otherwise incomplete combustion will take
place in the flash furnace. Excess air from 6&90% at the feed end is not
uncommon on such kilns. Through experience it has become known that
large percentages of excess air renders a kiln inefficient by lowering the
thermal level of the gases. As a whole, such a precalciner might be
efficient but the fact remains that the burning of the fuel in the burning
zone is done by large amounts of excess air hence no optimum conditions
prevail here. This, too, is the reason why a precalciner, without tertiary
air duct, usually does not achieve the high output rates that a precalciner of
equal size with tertiary air duct does.
Combustion air requirements to bum fuel (solid or liquid) can be
calculated as follows:

0 - (1 + i) [,.ll,,,, + 0.34783 &- $L)+ 0.04348AS ]

where:
o = combustion air required (kg air/kg fuel) - (lb air/lb fuel)
m = percent excess air in exit gas (use 5.0)
AC = percent carbon in fuel
AH = percent hydrogen in fuel
A0 = percent oxygen in fuel
As = percent sulfur in fuel
THE ROTARY CEMENT KILN

products of combustion are mathematically determined by the following


formulas:

CO2 from fuel = 0.0367AcW~ ...... (5-9)


SO2 from fuel = 0.02A~W~ ...... (5-10)
H20 from fuel = 0.0!3A~~W~ ...... (5-11)

N2 from fuel =
AN
- + 3.3478 ( 0.0267Ac + O.OIAs + 0.08AIi - O.OIAO)
100 . . . . . 1.
......
W,
(5-12)
Subtotal

Add excess air m (subtotal) . ...**


100
G,=tota.l . ..*** (5-13)

where:
G, = total combustion product (lb/lb clinker), (kg/kg clinker)
AC = percent carbon in fuel
As = percent sulfur in fuel
AH = percent hydrogen in fuel
AN = percent nitrogen in fuel
A0 = percent oxygen in fuel
m = percent excess air in exit gas (use 5.0)
WA = lb fuel fired/lb clinker or kg fuel/kg clinker

Carbon Dioxide in the Kiln Exit Gases.

The percent of oxygen in kiln exit gases is solely a function of the


amount of air introduced into the kiln and the amount of air consumed in
combustion of the fuel. In contrast to this, the percent of carbon dioxide
in the exit gases is influenced by calcination of the raw feed in the kiln as
well as by combustion.
When burning pulverized coal under perfect conditions, that is, when
no excess air is present in the kiln gases, a maximum of about 18.3%
COMBUSIION

CO2 would result from the combustion if there were no evolution of CO2
from the raw feed in the kiln. For fuel oil this value is about 15.8%, and
for natural gas, 12.2%. These percentages are obtainable only when per-
fect combustion conditions prevail in the kiln, which is very seldom the
case. Any excess air present will cause a reduction in the percent of CO,
in the gases. However, analysis of kiln exit gases in a cement plant
shows a percentage between 22 and 28% CO2 Therefore, a portion of the
CO2 in the kiln exit gases must originate from calcination of the raw feed.
Assuming that the CO, originating from the kiln feed remains
constant, then any drop in CO;! in the exit gases indicates an increase in air
in the kiln. Knowing that a rise in excess air results in poorer fuel
efficiency, it is then true that a decrease in CO, in the exit gases indicates
a higher heat consumption for each ton of clinker burned. Hence the
operator should try to obtain the maximum percentage of C$ possible
during normal operation of a kiln.
If the only function of the CO2 analyzer were to indicate fuel efficiency,
there would be no reason for its use as this information is available from
the oxygen analyzer. The conditions under which a carbon dioxide recorder
becomes of value is when calcining conditions change in the kiln, that is,
when the kiln is in an upset state. A change in the feed rate (including the
dust return rate, if any), feed composition, or feed advancement within the
kiln, can result in a change in the CO, content in the exit gases even at
constant combustion conditions in the burning zone. The reason why the
CO2 changes in these instances is because more feed, less feed, or different
feed, is being calcined in the kiln.
Changes that occur in the calcining zone cannot be seen by the kiln
operator when he looks into the kiln to observe the burning zone because
calcining takes place behind the burning zone. For this reason, a change
in the character of the feed bed could occur unknown to the operator until it
later becomes visible in the burning zone. Most common of such occur-
rences are kiln feed dust waves that flush at high speed into the burning
zone, the arrival of large chunks of scale, or a change in the amount of feed
entering the burning zone. By the use of a carbon dioxide analyzer and
recorder, the operator is able to recognize such changes before they become
visible in the burning zone, thus giving him time to make the necessary
adjustments in the kiln control variables sooner than he could without the
analyzer. If the operator notices an increase or decrease in the carbon
dioxide recording, no adjustments having been made to such variables as
fuel rate, air flow rate, or kiln speed, he knows that some change has taken
THE ROTARY CEMENT KILN

place in the calcining zone. This is an early warning of possible changes


in bumability that will soon occur. Once an operator has become
accustomed to a particular kiln, he will know what settings in CR, 02
fuel rate, feed rate, and other variables will most likely result in an effr-
cient and stable kiln operation for a given feed composition..
A carbon dioxide analyzer is also helpful in obtaining stable operating
conditions again after a kiln upset. Assuming that the kiln has been in an
upset condition with oxygen lower and CO2 higher than normal, then
when the recorders start to show an increase in oxygen, with the corre-
sponding decrease in CO2 without any change in feed rate, the operator
will know that calcining conditions behind the burning zone am improving
and he can expect an increase in burning zone temperature within a few
minutes.
Proper combustion requires that the kiln be operated in such a manner
thar

1. The kiln exit gases have an oxygen content of not less than 0.7%
nor more than 3.5% under normal operating conditions.
2. The kiln exit gases contain no carbon monoxide.
3. The kiln exit gases contain the maximum percentage of carbon
dioxide.

In addition, the operator should

4. Strive for optimum combustion conditions at all times by sta-


bilizing kiln gas oxygen content between 1.0 and 1.5%.
5. Take immediate steps to eliminate combustibles in the exit gases
whenever any carbon monoxide is detected therein.

In preheater and pmcalciner kilns it is of interest to note by how much


(percent) the kiln feed is calcined after it leaves the preheater vessels and
enters the rotary kiln proper. The laboratory checks this by analyzing the
feed before and after the preheater tower, then calculates the result as
follows:

% calcination = C--d
- loo (S-14)
C
COMBUSTION

where:
C= ignition loss of fresh kiln feed
d = ignition loss of feed after precalcination
In the second group are the variables over which the kiln operator has
control in order to obtain desired flame characteristics. These are:

Group 2
Fineness of the coal burned
Temperature of the fuel oil
Temperature of primary air
Temperature of secondary air
Flow rate of primary air
Flow rate of secondary air
Fuel rate
Burning zone wall temperature
Position of the burner
Degree of purity (dust concentration) of combustion air entering
the kiln
Cross-sectional loading of the kiln
Air requirement of the fuel
Density of the primary stream (primary air plus fuel)
Density of the combustion gases

The nine italicized variables are the ones indicated by Gygi* as being
the most important factors affecting the flame in a rotary kiln. A change
in flame characteristics will take place whenever one of the variables in
either group is changed. If the change results in a dangerous or undesirable
flame, one or more of the other variables must be adjusted to counteract
the bad flame condition.
Flame characteristics can vary considerably from one cement plant to
another, sometimes even from one kiln to another in the same plant The
reason for this is that a flame must always be tailored to existing kiln
designs and prevailing operating conditions. Clinker quality, refractory,
presence of rings, and kiln equipment problems force an operator to obtain
a certain flame that best fits the actual conditions in the particular kiln
under consideration.

*Gygi, H. “Warmetechnische Untersuchungen des Drehofens zur Herstellung v o n


Portlandzementklinker,” Dissemtion ETH, Zurich, 1937.
THE ROTARY CEMENT KILN

As a general rule in flame control, one should attempt to obtain the


shortest possible fire and the highest possible flame temperature without
adversely influencing clinker quality, coating formation, ring formation,
and refractory life, or causing damage to the kiln equipment in the dis-
charge area. Furthermore, the flame must not cause overheating in the
burner hood, kiln discharge end, or cooler. Once the ideal flame char-
acteristics have been obtained, the operator should make every effort to
operate the kiln in such a fashion as to cause a minimum of disturbance to
the flame. A flame should not willfully be changed during the course of
kiln operation unless specific conditions necessitate a change.
Now consider in detail the several flame characteristics that were men-
tioned at the beginning of this chapter, and how these characteristics can be
modified.
6.

The Flame

To an untrained eye, all flames in the burning zone of a rotary kiln


appear more or less the same. As shown in Fig. 6.1, there is a short
plume of air and fuel emanating from the nozzle, then at the end of the
plume the fuel ignites and forms the flame. Flames vary considerably in
length, shape, color, direction, and point of ignition. With time the
novice learns to recognize these characteristics, their effect on kiln effi-
ciency, and how to control them. Such terms as short, long, lazy, snappy,
cold flame, and hot flame, the terms used to describe the different shapes
and colors of flames, take on a real meaning and soon become com-
monplace in the language of the kiln operator.
The burner end of a kiln is equipped with a port through which the
operator can view the burning zone. Because of the brilliancy of the incan-
descent fuel and gas, it is necessary to look through a special colored glass
to protect the operator’s eyes. When observing the flame, the operator
must judge the point of ignition of the fuel, length, direction, and shape of
the flame, and the color of the flame, the color being a good indication of
flame temperature.

6.1 FLAME CHARACTERISTICS

Under stable operating conditions, no changes to the burner assembly or


position are normally made unless hot kiln shell conditions make such an
adjustment necessary. Therefore, the flame characteristics should theo-
THE ROTARY CEMENT KILN

retically remain unchanged. This, however, is not the case, because flame
shape is influenced by many factors such as I.D. fan speed, secondary air
temperature, primary air pressure and temperature, as well as changing
conditions in the burning zone itself.
The factors that influence these characteristics can be subdivided into
two groups. In the first group are the variables over which a kiln operator
has little or no control. Some of these variables are impossible to change
because they are an integral part of the kiln itself. Those that could possi-
bly be adjusted, require either the help of a third party or a kiln shutdown.
The following listed elements can be included in this group:

Group I
Diameter of primary air nozzle
Orifice size of fuel burner
Diameter of the kiln
Design of the burner
Heat value of the fuel

6.2 LENGTH OF THE FLAME

When considering the length of a flame, one must make a clear dis-
tinction between two aspects of flame length. Flame length can be referred
to as the distance between the nozzle of the burner and the end of the
flame, or it can be expressed as the distance between the point where igni-
tion of the fuel starts and where the reaction process of fuel combustion
ends, that is, the length of the ignited part of the flame. The difference
between these two concepts is often overlooked, but it is important to an
understanding of kiln operation. Fig. 6.2 shows clearly the difference be-
tween these two measurements. Comparing the two flames A and B, it is
seen that the distance between the nozzle of the burner and the end of the
flame, 44 ft, is identical on both flames. If the flame length is considered
as the ignited part of the flame only, then flame A is 40 ft long and flame
B is 28 ft. In this book, the terms total flame length and ignited flame
length will be used to distinguish the two measurements.
The variables having the greatest influence on flame length are the per-
centage of combustion air present and the velocity of the fuel-air mixture
at the tip of the burner.
THE FLAME

PRIMARY AIR PIPE

BURNER

Fig. 6.1 The flames from all types of fuels have certain common
characteristics.

It was seen in the previous section on combustion how a fuel, when


introduced into the kiln, must combine with oxygen to start the chemical
reaction of combustion. In other words, air which contains the oxygen,
and fuel have to come into contact with each other to make the fuel burn.
A lack of air can cause the flame to become too long, as the fuel then has
to search for oxygen further back in the kiln.
Burner-tip velocity is the primary influence on flame configuration
when coal and/or coke are fired. Unfortunately, direct-fued kilns most
often demand a burner orifice size that is not to the best advantage of good
flame control. On such systems, the burner size is more often governed
by the coal mill’s minimum air evacuation requirements and coal mill fan
static pressure limitations. Flame control, unfortunately, is of secondary
priority on such systems. Recently new, more sophisticated coal burner
designs have come into use to eliminate some of the shortcomings of
direct-fired systems but they still present problems in flame-control cap-
ability. It is known that a large percentage of indirect-fired kilns, par-
ticularly in Europe, operate with tip velocities of between 70-83 m/s
(14,000-16,500 ft/min) and show exceptionally good flame patterns.
However, direct-fired kilns more often are forced, by system &sign, to
operate with velocities of between 45-66 m/s (9,000-13,000 ft/min)
which mostly deliver a relatively lazy swirling flame. Many operators on
such direct-fired kilns have unsuccessfully attempted to reduce the burner
orifice size (i.e., to increase the tip velocity) and experienced coal evacua-
tion problems from the mill. In other cases, the combination of higher tip
velocity and relatively cold primary air have caused unstable and delayed
ignition points in the flame. Coal burner &signs on direct-fired kilns
must be viewed as a compromise solution and is one area that leaves a lot
of room for improvements.
THE ROTARY CEMENT KILN

Operators should be told by the engineering department what the typical


tip velocity is on their kiln under various primary air pressures. As a rule
of thumb, it takes a minimum of 35-m/s (7,000-ft/min) velocity in cir-
cular pipes to keep fmely ground coal particles in suspension. Below this
minimum velocity, coal could settle in the pipe and form pockets that
could lead to dangerous backfines and/or explosions. Consult with the
plant engineer about the minimum permissible, safe primary air pressure
and once this level is established, never introduce coal into the
system unless you are absolutely sure that the air velocity
and pressure are above this minimum permissible level.

6.3 FLAME PROPAGATION SPEED

For coal-fired kilns, the primary air velocity should be at least twice as
high as the flame propagation speed to prevent flashbacks of the flame.
Flame propagation is usually considerably lower than the velocity needed
to convey coal dust by means of primary air into the kiln. Therefore, the
minimum velocity necessary to convey coal without settling in ducts takes
precedence over flame propagation speed when setting air-flow rates or
designing new burners.
W. Ruhlande in his investigation of flames, found that a decrease in
nozzle diameter will at first give a shorter flame, but further decreases in
diameter could lead to longer flames. He also found that no change in
flame length will take place when the ratio between total combustion air
and fuel introduced into the kiln remains constant. From this it can be
concluded that the flame length is maintained unchanged when the percent
oxygen in the exit gases remains the same. In other words, each time the
fuel rate is changed, the rate of air going to the kiln should also be changed
in order to maintain the same flame length.
The total mass of combustion air entering the kiln is the sum of the
primary air, the secondary air, and the so-called parasite air which enters
the kiln through leaks at the kiln discharge and burner hood. Because the
operator has no control over the last air flow (unless he forgets to close
one of the large doors of the burner hood), we shall consider only the fust
two mass flows.
The amount of total combustion air entering the kiln is governed
mainly by the speed of the induced draft fan (usually called the I.D. fan). If
THE FLAME

TABLE 6.1
FLAME PROPAGATION

4 (Metric Units)

Primary air Flame propagation (m/s)


m3/kgcoal 30% vM* 30% VM 10% VM
5% ash 15% ash 7% ash

1 4.0 3.9 2.2


2 8.8 7.5 3.8
3 13.0 11.0 5.4
4 14.4 11.7 6.7
14.0 11.1 6.3
6” 13.1 10.0 5.5
7 12.2 9.0 4.9
8 11.4 8.1 4.5
9 10.8 7.5 4.1
10 10.3 6.9 3.8
11 9.8 6.6 3.6

4 (English Units)

Primary air Flame propagation (ftlmin)


ft3/lb coal 30% VM 30% VM lO%VM
5% ash 15% ash 7% ash

20 950 900 500


40 2250 1900 900
60 2800 2300 1300
80 2750 2180 1240
100 2550 1900 1050
120 2320 1670 930
140 2150 1500 840
160 2020 1350 750
180 1900 1250 690

* volatile matter
THE ROTARY CEMENT KILN

the fan speed is increased with the fuel rate constant, more combustion air
enters the kiln. Conversely, a decrease in ID. fan speed will result in a
decrease in total combustion air entering the kiln. The operator has at his
disposal a quick indication with the kiln exit gas oxygen analyzer to deter-
mine if the flame theoretically is getting longer or shorter. Lack of com-
bustion air in the kiln is indicated when the oxygen analyzer reading ap-
proaches zero and the carbon monoxide combustible instrument shows that
carbon monoxide is present in the kiln exit gases.
To explain this effect on flame length consider several examples.

Case 1: Oxygen reading is 2.0%, ID. fan speed is increased such


that the oxygen reaches a value of 4.0%. Result: total
flame has been lengthened.
Case 2: Oxygen reading is 2.0%, ID. fan speed is decreased such
that oxygen decreases to a value of 0.7%. Result: total
flame length has been shortened
Case 3: Oxygen reading is 0.7%, I.D. fan speed is decreased such
that the oxygen obtains a value of 0.1% and a small
amount of CO is also indicated. Result: total flame has
been lengthened.
Case 4: Oxygen reading is 0.1% and combustibles are showing.
I.D. fan speed is increased such that the oxygen increases
to 0.7% and no more combustibles are indicated. Result:
total flame has been shortened.

From these examples it becomes clear that it is not possible to state


simply that an increase in air flow will result in a lengthening of the flame
or a decrease will shorten the flame, but that the resulting changes are
dependent on the percent oxygen present before the changes in air-flow
rates are made.
The total flame length could become infinitely long under the condition
of an extreme lack of air. As a matter of fact the fuel could travel the
entire length of the kiln without igniting because not enough air is present
in the kiln itself. Upon entering the dust collector at the kiln rear where
plenty of air is available, the fuel could explode if the temperature is high
enough to ignite the fuel. One can easily visualize the results of such a
disastrous occurrence.
A word of caution must be expressed at this point Considering how
flame length can be changed by changes in the I.D. fan Speed, and knowing
THE FLAME

the desirability of maintaining the flame length constant, could lead one to
the conclusion that the I.D. fan speed should be proportionally changed
whenever a change in the fuel rate is carried out. Later on in this book,
the pitfalls of such an approach because of its effect on back-end
temperature and overall kiln operation in general will be pointed out. One
must at all times correlate the different control functions together for the
entire kiln operation and take the approaches that will be the most
beneficial for stability of kiln operation. For example, maintaining the
oxygen content at a fairly constant level while varying the kiln speed is
one technique of kiln burning that is not recommended for prolonged stable
kiln operation.
In the technique of kiln burning stressed in this handbook, the operator
gives special attention to the burning zone temperature and back-end tem-
perature, at the same time maintaining the oxygen content in the kiln exit
gases within a range of not less than 0.7% and not more than 3.5%. This
“floating” control of the oxygen results in changes of the total flame
length. Experience has shown that the flame length changes are of minor
magnitude when the oxygen level is controlled within this range, and far
more benefits can be derived in operating stability than when the oxygen
would be strictly controlled at a constant level.

6.4 IGNITION OF THE FUEL

The initial ignition (start of combustion) of the fuel is primarily depen-


dent on sufficient heat to ignite the fuel and on sufficient air to obtain
combustion of the fuel. A deficiency in either one of the two in the area
where the fuel leaves the burner and enters the kiln can cause a delayed
ignition of the fuel.
During the initial period of a kiln start, the kiln operator may ex-
perience difficulties in obtaining or maintaining continuous ignition of the
fuel. This is understandable because during such periods furnace tempera-
tures are usually too low to ignite the fuel properly. One must resort to
the help of a pilot burner (auxiliary torch) placed at the mouth of the
burner pipe to obtain good ignition. It is a good practice to leave the pilot
burner in this position until the kiln interior has acquired the necessary
heat to sustain continuous ignition. Operator jargon often refers to this as
the time when the wall temperature in the burning zone can support the
flame.
THE ROTARY CEMENT KILN

Once the kiln has reached operating temperature and continuous burning
of the fuel has been obtained, the ignition point of the flame can be will-
fully adjusted or inadvertently changed by changes of certain operating
variables.
A decrease either in primary or secondary air temperature can move the
ignition point further into the kiln. Design of the kiln burner hood and
the burner also plays a part in the point of ignition of the fuel. For ex-
ample, the secondary air stream could enter the kiln in such a fashion that
a rapid contact with the fuel is not possible. The fuel will thus ignite at a
point further in the burning zone, or a burner can be designed so that a
rapid mixing of the air with the fuel is promoted which results in an earlier
ignition of the fuel once it enters the kiln.
It must be pointed out that no clear-cut general answer can be given to
the question of where the point of fuel ignition shoud be located in a rotary
kiln. This depends on the type of fuel used and the overall conditions that
result from a given flame structure and ignition point location. Although
it is advantageous to have the fuel ignite as early as possible after it leaves
the burner, this can in many cases create overheated conditions in the kiln
nose area and the burner hood, as well as in the cooler. In general, one can
say that the earliest possible ignition point of the fuel is desirable under
the condition that there will be no harmful effects on kiln equipment as a
result.
The plume, as shown in Fig. 6.1, is the part of the flame between the
burner nozzle and the point of ignition of the fuel. On coal-and-oil-fired
kiln this plume is recognizable as a black cloud or jet. On coal-fired kilns,
a shorter plume (earlier ignition) can be obtained by grinding the coal to a
larger surface area; that is, grinding the coal finer. Earlier ignition on oil-
fued kilns can be accomplished by increasing the temperature of the fuel
oil, and by selecting a smaller oil burner opening (orifice, or tip size) such
that the oil will be better atomized. For both types of fuels, shorter
plumes can be obtained by designing the burner so that a rapid mixing of
air and fuel takes place at the burner tip. Control of plume length in a gas-
fired kiln is not easily accomplished, but is not particularly important.
Kiln interior temperature, primary air temperature, and secondary air
temperature, three very important factors affecting plume length, are the
variables that are most frequently apt to change during operation of the
kiln. An operator will always try to hold these temperatures within close
limits, because a large change in any one can lead the kiln into an upset
condition and affect overall flame characteristics in an undesirable manner.
THE FUME

At this point, a word of caution must be inserted with respect to the


primary air temperature. On many rotary kilns, preheated primary air is
used to promote ignition of the fuel and to form a specific desirable flame.
The ignition temperature for various fuels commonly used on rotary kilns
can be as low as 800 F (425 C). For safety reasons, temperatures within
the primary air duct must therefore be kept well below this critical level at
all times. This becomes especially important on coal-fired kilns where the
coal dust is mixed with the primary air inside the burner pipe. Exceeding
this critical temperature could ignite the coal inside the burner pipe with
possible disastrous results. The most probable times for this adverse pos-
sibility are the periods immediately after a fire has been cut off or lighted
during short kiln shutdowns; that is, when the burning zone is still at an
elevated temperature. These high burning-zone temperatures could easily
find their way into the primary air duct. Explosions have occurred in the
primary air pipe because fuel was able to accumulate inside the pipe during
a shutdown as a result of a leak in the coal feeder mechanism or leak of the
fuel shut-off valve. Because of the potential danger created by such
mechanical deficiencies, these fuel valves and feeders should be inspected
frequently on a routine basis and any malfunction repaired at the earliest
possible time. To safeguard against an explosion due to ill-adjusted air-to-
fuel ratio during the initial moment of firing a kiln, an operator will
always make sure that enough air is present in the kiln for the intended
fuel rate and that enough heat is present to ignite the fuel before the fuel is
introduced into the kiln.
So far, we have centered our attention on the ignition point and the
plume of the flame, giving us some insight into one important aspect of
ignition. Now let us consider another important element, the aspect of
total ignition of all the fuel introduced into the kiln.
It is absolutely essential that all the combustion reaction (burning of
the fuel) takes place in the atmosphere of the burning zone. Impingement
of partially burned fuel upon the feed bed or kiln wall must be avoided at
all times when a kiln is being fired with coal or oil. Impingement of un-
burned fuel particles on the wall and feed bed is not only harmful to the
refractory and clinker but also acts as a bad influence on fuel efficiency and
kiln operating stability. Alignment of the burner, fineness of the coal, and
the velocity with which the fuel and the primary air enter the kiln am the
most important factors governing impingement. Fineness of the coal
must be such that all coal particles are burned completely while they are in
suspension in the kiln atmosphere. The same holds true for fuel oil firing;
THE ROTARY CEMENT KILN

oil droplets must be small enough so they bum completely in the same
manner. It is obvious that the velocity of the fuel as it enters the kiln
must be adjusted properly so the fuel particles have enough velocity to
remain in suspension while they bum, but not so high that they are carried
too far into the kiln. A wrong selection of primary air pipe nozzle diam-
eter or orifice size of oil and gas burner nozzle could result in either insuf-
ficient or excessive velocity of the fuel-primary air mixture.
All coal-mill systems are equipped with thermocouples to measure the
so-called coal-mill outlet temperature. The maximum permissible and
recommended temperature is 74 C (165 F) for high-volat& coal and 85 C
(185 F) for low-volatile coal. This is considerably below the ignition
temperature for these types of coal but considered essential as a safety
measure to prevent premature ignition in the mill or primary air pipe.
There is another control function an operator has to keep close watch
over, that is, the coal-mill inlet temperature. The coal mill’s function is
not only to grind and pulverize but also to dry the coal. ,This is accom-
plished by using hot excess air from the clinker cooler, cooler air which
can fluctuate quite frequently from a high of 650 C (1202 F) to a low of
200 C (393 F). From the previous discussion of the combustion triangle.
(Fuel-Air-Heat), it can be seen that such high air temperatures would be
sufficient to start a fire in the coal mill if this hot air temperature is
allowed to fluctuate unchecked. For this reason, the hot air duct to the
coal mill is equipped with a safety ambient-air inlet device that allows, by
means of a damper, the mixture of hot with cold air, and thus controls the
coal-mill temperature within safe limits. Coal, too, fed to the mill for
grinding and drying, can fluctuate considerably in its requirement for heat
depending on its moisture content and the amount being fed to the mill.
Therefore, it becomes apparent that here three highly varying entities
combine together to form one of the seemingly more complicated control
functions on the kiln system:

changing cooler air temperatures


changing coal moistures
changing coal feed rates

This process control function is handled with a few simple instruments.


First, the designated main control (input) variable is the coal-mill outlet
temperature from a thermocouple located near the point where the coal/air
mixture leaves the mill. The operator regulates the setpoint for this tem-
THE FLAh4E

perature on the automatic controller, which in turn will almost con-


tinuously adjust the ambient air-tempering damper in the hot inlet-air duct
to maintain the desired mill outlet temperature at this predetermined set-
point. In the case where the cooler excess air is so hot that the ambient air
damper, even if wide open, is not sufficient to hold the coal mill outlet
temperature below safe levels, there is another control loop which can be
found on most coal mill systems. Here the input variable is the coal mill
inlet temperature. Here, too, the operator sets a predetermined setpoint
for this temperature on a separate controller which in turn drives a damper
that allows for ambient air mixing at a second location (usually at the
beginning of the hot-air duct near the cooler). The newer, more modem
sytems also incorporate automatic CO, injection as an added safety device
to prevent coal-mill fires and/or explosions. If properly designed, operated,
and maintained, these systems are safe. There are, however, some impor-
tant aspects and considerations an operator must always keep in mind
regardless of how well or calm the system might appear. It is important
to regard it with caution and expect the unexpected. Here are a few
pointers:

a) Both of the above-mentioned controllers must be set so that they


respond very rapidly to a slight change in temperature, i.e., the
tempering damper adjustment must be almost instantaneous
whenever a slight deviation from setpoint occurs.
b) All input-output devices must be properly maintained, they must
function, and must be immediately attended to if adjustments are
needed. No shortcuts can be taken.
c) Under normal operating conditions, both ambient-air-tempering
dampers must be set in a position so that there is plenty of lee-
way (adjustment capability) left for the damper to move in either
direction. Case in point it wouldn’t do a controller any good if
the tempering damper, under normal operating conditions, is al-
ways in the fully open position.
d) It is important to communicate with the plant engineer so as to
get a clear understanding of the maximum and minimum per-
missible temperatures and pressures. A record of these tempera-
tures should be kept in a standard operating-procedure notebook
e) Learning the coal-mill system should be the first priority in the
learning program. Becoming fully familiar with all its aspects
THE ROTARY CEMENT KILN

and learning what to do when things go wrong are essential. Just


learning what buttons to push for what isn’t enough.

Having started this discussion with flame control and arrived at a dis-
cussion of where hot air is leaving the cooler, demonstrates how inter-
related the functions of a rotary cement kiln are. A change in one variable
can ultimately cause a change at another location that, on the surface,
doesn’t appear to be related at all.
Returning to the discussion of flame control, the theoretical flame tem-
perature can be approximated by the following formula:

English or Metric units:

T=
H” (6-U
l.lL4s

where.:
T = theoretical flame temperature (“C or OF)
k = combustion air required (lb/lb or kg/kg of fuel)
Hv = heating value of fuel @al/kg or BtuIlb)
S = specific heat of combustion gas (use 0.29)

The values obtained by these calculations are usually in the magnitude of


3700-4800 F (2040-2650 C). At these temperatures the flame should
have a dazzling white color. Comparing this with the color usually ob-
served in a rotary kiln, it becomes quite obvious that actual flame tempera-
tures fall quite short of these calculated maximum possible temperatures.
Flames with a light yellow to white color [24OCM300 F (1320-1540 C)]
and yellow to light yellow [2000-2400 F (1090-1320 C)] are common in
rotary kilns. The flame colors discussed here are actual flame color and not
the color an operator sees through a filter glass.
Operators often speak of hot and cold flames. These terms, although
contradictory because there is no such thing as a “cold” flame, are fully ac-
ceptable because they state in unmistakable terms a given characteristic of
a flame. What the operators mean by these expressions are high and low
flame temperatures.
THE FLAME

6.5 OXYGEN ENRICHMENT

The steel industry has made some large improvements in regard to


flame temperatures in open-hearth furnaces by oxygen enrichment of the
combustion air, resulting in much higher flame temperatures, which in
turn improves the efficiency of this type of furnace. Oxygen enrichment
of the combustion gases in a rotary cement kiln has been attempted in a
number of cement mills, but because of the high cost the process has been
abandoned. Using oxygen enrichment could create an undesirable result
because of the extremely intense, hot, and short fire that is characteristic of
this type of flame. Extreme high temperatures might be localized close to
the discharge area of the kiln which could cause damage to the refractory as
well as to the burner hood and cooler.
It is quite conceivable that oxygen enrichment might, once again, find
some favor and acceptance in a precalciner kiln. To the author’s knowl-
edge no plant has tried this so far. As previously stated, there are two
types of precalciner kilns: the ones with tertiary air ducts and those that
draw the combustion air for the flash furnace from the kiln itself. It has
been mentioned that the latter must operate at very high kiln exit-gas
oxygen levels to secure the proper air supply to the flash furnace. This
leads to lower flame temperatures in the kiln itself due to the excess air
present but makes this type of kiln considerably less costly than a
pmcalciner with a tertiary air duct. Besides, control of the kiln overall is
also made easier by not having to control the air flow and temperature
through the tertiary duct. Oxygen enrichment at the flash furnace could
possibly allow such kilns, without a tertiary air duct, to operate the kiln at
normal oxygen levels in the kiln exit gas of say O&1.5%. This idea
would require research before deciding whether or not it could be feasibly
implemented.
Flame D, Fig. 6.2, shows an extreme condition in which part of the
flame impacts against a clinker ring, a common condition on some kilns.
Although such a ring formation speeds up the reaction of combustion (if
enough air is present), this condition is extremely harmful to the coating
and refractory in the area immediately preceding the ring. It is not uncom-
mon, on kilns that experience such frequent ring problems, to find the
shortest refractory life of the entire burning zone lining in this particular
area where flame impingement upon the wall takes place. This is one of
the masons why it is a good practice to remove (shoot) the ring before it
causes flame impingement, that is, before the ring has grown too large.
THE ROTARY CEMENT KILN

b
A
I

OFT

Fig. 6.2 In an exaggerated and simplified form, the illustrated flames show
the essential differences between the several flames.
THE FLAME

6.6 SHAPE OF THE FLAME

As previously pointed out, it is desirable to operate a rotary kiln with


the total flame length as short as possible. A short flame extends the
length of the calcining zone, which can often lead to increased production
capabilities of the kiln.
Now consider Flames A and C in Fig. 6.2. Although the same amount
of fuel is being burned in both flames, the two flames are entirely differ-
ent. Flame A represents what operators call a “lazy” fire because heat from
this flame is released over a relatively long distance of the burning zone.
Flame C is a “snappy” fire in which the heat is released over a shorter
length of burning zone. Both types of flames can be found in rotary kilns,
but flame C is more favorable for ease of kiln control, clinker quality, and
fuel efficiency.
Flames such as shown in example A usually have a tendency to float
about the kiln atmosphere in an unstable fashion, but flame C, with its so-
called better body, tends to remain fixed in position better when small
changes in air-flow rates or temperature occur. Flame C, with only a few
exemptions, is the most favorably shaped flame for a rotary kiln. The
burning fuel (flame) expands in a uniform wedge equally in all directions
along the kiln axis, and is not too short and not too long in total length.
Stability of the flame must always be majlitained and stable flame charac-
teristics (when the flame is favorably formed) pay off with long refractory
life and good kiln control conditions.

6.7 DIRECTION OF THE FLAME

One of the most important things for the kiln operator to remember is
the fact that the path of the flame is not a straight line, and it does not
remain constant at all times. Thus, when the primary air pipe is set for
the fuel-primary air mixture to travel toward a certain target in the burning
zone, this will not guarantee that the flame itself will follow this par-
ticular path for its entire length. The flame has a tendency to lift upward
toward the top of the kiln arch, because of the buoyancy resulting from
uneven entrance of secondary air into the kiln. This hot secondary air
stream enters the kiln from the bottom, and moves upward in a curved path
THE ROTARY CEMENT KILN

toward the axis of the flame a few feet past the burner nozzle, where it
starts to mix with the fuel-primary air stream.
Another factor influencing flame direction is the mechanical condition
of the primary air pipe nozzle. Because of the high temperatures pre-
vailing under normal operating conditions in the burner hood area, this
primary air pipe nozzle is especially vulnerable to warping and distortion,
resulting in uneven envelopment of primary air around the fuel jet causing
erratic flame shapes and direction.

6.8 ADJUSTMENT OF FLAME DIRECTION

On some rotary kilns, the primary air pipe is fixed slightly below the
center of the kiln, in an effort to offset the flame buoyancy. Most kilns,
however, have facilities for adjusting the position of the primary air pipe
and the burners, thus the operator can change flame direction and, to a
limited degree, change the shape of the flame. Every adjustment in the pri-
mary air pipe or burner position will bring about a change in flame char-
acteristics and can therefore affect burning conditions in the burning zone.
Buoyancy of the flame can usually be counteracted by a slight downward
tilt of the primary air pipe so the main body of the flame will be centered
at a certain distance inside the burning zone. In Fig. 6.2, flame E shows a
buoyant flame with the primary air pipe centered in the kiln axis. By
nltmg the pipe down slightly, the flame assumes the position shown by
flame F.
Because so many factors influence flame direction, it is necessary to fix
the primary air pipe position according to the actual flame condition en-
countered in each individual kiln. To do this, the operator will observe the
flame in the kiln, paying particular attention to the position of the flame
body at point “x” in Fig. 6.2. Fig. 6.3 is a cross section of the kiln at
‘LX.))
The centerpoint of the flame at this location can be in any one of the
indicated squares depending on how well the flame direction is adjusted. It
is generally assumed that the best heat exchange between the flame and the
feed takes place when the flame is pointed slightly toward the feed bed,
hence the most favorable target for the flame is position 2A or on the kiln
axis (2B). On the other hand, if the flame is directed too close to the feed
bed (3A), there is danger that part of unburned fuel (especially on coal and
THE FLAME

Fig. 6.3 Imaginary targets in the kiln indicate the ColTed and incorrect areas
toward which the flame should be directed.

oil fires) could enter the feed bed, a condition that is highly undesirable. A
flame target leaning too much toward the kiln wall (lC, 2C, or 1B) could
result in flame impingement upon the coating, an action which is known
to shorten the life of the refractory.
It has been found that the best method to fix the burner position is in
the following manner: Prior to firing up a kiln (after a relining job),
measure a given distance, e.g., 20 m (65 ft) from the kiln discharge and lay
a white rag at this location. Then, go back to the firing floor and look
through the burner (a cross-hair at the tip is helpful) to target the burner
direction toward this rag. Once this is accomplished, tilt the burner slight-
THE ROTARY CEMENT KILN

ly toward the load-side, e.g., the 5- or 7-o’clock position from the rag).
The distance to place this rag is found by trial and error, for in some
instances 80 ft is required and other times only 50 ft is necessary to obtain
optimum flame positioning. Once this position has been found, however,
it should not be changed and should become standard procedure on all
subsequent kiln-start preparations.
Although no clear-cut standards can be given for the direction of the
flame for all kilns, because each kiln has this position specifically tailored
to its own particular problems, design, or conditon, there are, however, a
few rules that can be applied to all flames, regardless of what rotary kiln is
under consideration. These are:

a) When the primary air pipe nozzle has accidentally been warped,
resulting in an erratic flame shape and direction, immediate steps
should be taken to repair this condition.
b) A flame should never be allowed to impinge upon the coating or
bare refractory for a prolonged length of time.
c) A flame should never be allowed to strike too hard upon the feed

d) Oil burners or gas burners should be centered well in the primary


air pipe in order that an even envelopment of air around the fuel
jet takes place.
e) Flame direction should be adjusted only when the kiln is in stable
operating condition and the temperatures, fuel pressures, and air-
flow rates are at a normal level. Flame direction changes can be
caused by unusual operating conditions. If an attempt were made
to adjust the flame direction at such times, there will most likely
be an undesirable flame once the kiln returns to normal operating
conditions again.
f) It is better to make the desired adjustments in flame direction in
several small steps instead of a large one in order that the operat-
ing stability of the kiln is not affected adversely.
g) Once the ideal flame direction has been obtained, the primary air
pipe position should not be changed unless a definite reason (such
as to combat ring formation or hot shell conditions) makes it
desirable.
h) To protect the primary air pipe from possible damage during a
shutdown a certain amount of primary air flow must be main-
tained until the temperature inside the kiln is low enough (ap-
THE FLAME

proximately 600 F or 315 C) that the pipe cannot be damaged.


Upon power failure when the primary air fan stops, the primary
air pipe must be immediately removed from the burner hood.

6.9 TEMPERATURE OF THE FLAME

The amount of heat released by the flame depends on the flame temper-
ature, and the flame temperature indicated by the color of the flame (Table
6.2). By observing the color, the operator can estimate the flame tem-
perature within certain rather broad limits. However, before discussing
color of the flame, it is first necessary to point out that an increase in fuel
rate does not always increase heat output of the flame, nor does a fuel rate
decrease always reduce the heat output. The reason is that combustion of
the fuel is also dependent on the proper amount of air (oxygen) available
and upon the prevailing temperatures of the gases and the kiln wall in the
burning zone. Liberation of heat by the fuel, because of its dependence
upon these factors, can at times be somewhat erratic and, confusing to the
newcomer who is learning to operate a kiln.
A prime example, and because of its simplicity and frequency of occur-
rence, the best worth mentioning, is the condition in which the kiln is
operating with a very low content of oxygen in the exit gases, e.g., 0.5%.
The operator notices that the burning zone temperature starts to decrease.
The wrong procedure in many such instances is to increase the fuel rate,
hoping that this will help to raise the temperature. The contrary is most
likely to happen; that is, a further decrease in flame temperature, because
of the insufficient air supply available for good combustion of the fuel.
The proper procedure in such an instance would be to reduce the kiln speed
to allow more time for the clinker to bum properly, or to reduce the fuel
rate to improve combustion of the fuel. A third possibility is to increase
the primary air flow to compensate for the lack of necessary combustion
air brought about by the increase in fuel rate.
The same or similar reaction can take place with a sudden large drop in
secondary air temperature, or when too much air (large percentage of excess
air) is introduced into the kiln by excessively high I.D. fan speed. Both of
these occurrences can result in a decrease in flame temperature.
Under any operating condition, regardless of the fuel rate, it is desirable
to obtain the highest possible flame temperature. In other words, the best
THE ROTARY CEMENT KILN

fires in a rotary kiln are the ones with a very bright color. Orange and red-
colored flames are undesirable because they possess a lower flame tem-
perature. The values shown in Table 6.2 are not only helpful in respect to
flame temperatures, but can also be of service to the operator for any kind
of observation in the kiln system where elevated temperatures prevail such
as the burning zone wall, feed bed and kiln shell.

TABLE 6.2

CORRESPONDING TEMPERATURES
FOR OBSERVED COLORS

“F T

Lowest visible red 875 475


Lowest visible red to dark red 875-1200 475-650
Dark red to cherry red 1200-1375 650-750
Cherry red to bright the red 1375-1500 750-825
Bright cherry red to 1500-1650 825-900
Orange to yellow 1650-2000 900-1090
Yellow to light yellow 2000-2400 1090-1320
Light yellow to white 240&2800 1320-1540
White to dazzling white over 2800 over 1540

Of specific interest for the kiln operator are the following factors that
serve to raise flame temperature:

a) Increasing the secondary air temperature.


b) Using less primary air, thus making it possible to utilize more
secondary air which is preheated to higher temperature.
c) Promoting rapid mixing of the air and fuel upon leaving the
burner by improving the &sign of primary air pipe and burner.
d) Better atomization of the fuel oil by increasing the fuel oil
temperature and selecting a smaller burner orifice size (tip size),
or employing a mechanical device in the burner nozzle to bring
about better atomization.
e) Operating a kiln with neither a deficiency nor excess of air by
maintaining oxygen content of not less than 0.7% and not more
than 3.5%.
7.

Heat Transfer

In this chapter a review of the fundamental concepts of heat transfer is


presented The second law of thermodynamics states that heat always
flows from a point of higher temperature to /the point of lower tem-
perature. Heat can be transferred by radiation, which is the transfer of heat
from one body to another by means of heat waves without the two bodies
being in actual contact with each other. Another type of heat transfer is by
co~action, where the vibration of molecules of solid bodies collide with
adjacent molecules and thus transfer energy between each other while in
contact The third type of heat transfer is known as convection which
occurs in gaseous and liquid fluids. In this type of heat transfer, the mole-
cules, after having collided with each other, are free to move and circulate
and thus are able to transmit more energy to other molecules having differ-
ent temperatures.
In a rotary kiln, the largest amount of heat transfer occurs by radiation
and ConductiOll whereas heat transfer by convection occurs only in small
mounts. Although there are many more occurrences of heat transfer, the
following examples explain some of the more prominent heat exchanges
hat t&e place in a rotary kiln:

RADIATION: Flame + feed bed


Hot kiln wall ) feed bed
CONIXJCTION: Heat from kiln interior - kiln shell
Kiln chains ,feed bed
cO~ECTION (in heat exchangers):
Super-heated stem oil in tubular pipes
Hot kiln exit gases __) steam in boiler tubes
THE ROTARY CEMENT KILN

Kiln burning can be defined as exercising control over all heat transfers
that take place within and in the proximity of the rotary kiln. A kiln
operator has little control over the type of heat exchange that takes place.
Whether it will be radiation, conduction, or convection is primarily fmed
and governed by the design of the equipment. But control over the mount
of heat transfer that takes place is posible since the operator’s primary
function is to control the temperatures at multiple locations of the kiln
system. Temperature and heat transfer are quantities that are dependent on
each other and inseparable.

7.1 THE THERMAL WORK REQUIRED


IN A CEMENT KILN
Regardless of what type of kiln is used, there is a fMed amount of
thermal work that has to be done within the system to obtain what is
known as cement clinker. This them&al process is accomplished in several
stages, stages which are discussed in more detail in a later chapter. The
key thinuo remember is that there must be a minimum amount of heat
(temperature) imparted into the feed so that these phases can progress prop-
erly. WL. De Key&, based on laboratory investigations, has made deter-
minations as to the temperatures at which these phases occur. Likewise,
G. Martin2 has done extensive studies of the thermodynamics of a rotary
kiln, however, he based his investigations on practical kiln data. From
these two works the following table has been compiled which informs in
very broad terms what actually goes on inside a rotary kiln.
The concept of low-grade and high-grade heat, first advanced by G. Mar-
tin, is a very important aspect of kiln operation. It simply means that
regardless of how much heat is applied to the rear of the kiln, if that heat
is below 805 C (1481 F) there can be no other thermal work done other
than drying the feed and driving off the chemically combined water from
the clay minerals. No calcination can take place. The same holds true in
the burning zone. Unless a minimum of 1400 C (2552 F) is reached for
the feed in the burning zone, there will simply be no clinker produced that
could meet the quality standards.

7.2 HEAT PROFILE

Temperatures differ throughout the kiln and a very important factor is


the temperature difference between the material and the gases at any given
point in the kiln. A recorder chart of the gas temperature shows gas tem-
HEAT TRANSFER

TABLE 7.1
PHASE FORMATIONS

Thermal work Minimum temp.

“C “F
1. Evaporation of free water from 100 212 Required low-
the feed grade heat
2. Evolution of the chemically com- 550 1022 363 kcal&g
bined water in the clay minerats (652 Btu/lb)

1
3. Evolution of CO, from calcium
and magnesium carbonate
(calcination) 805 1481
4. Formation of interm phase C,F 800 1472
5. Formation of interim phase CF 900 1652
6. Formation of interim phase 1000 1832
‘2,s + CIAS + CF + C2F Required high-
7. Formation of interim phase 1 heat grade heat
C$ -I- CsAs + CsA + C.$ + 511 kcahkg
CF + C2F (919 Btu/lb)
8. Formation of interim phase
C2S + CsA + CSA3 + C2F +
c3s
9. Formation of C3A, C3S, C!,S, 1300 2372

10. ~ilibrium in CaA, CaS, 1400 2552


C,S + CsA2F

perature only, and this temperature never corresponds to the temperature of


the material in the same location. Also, conditions could change in such a
way that the gas temperatures would show a considerable change but the
material temperature remains unchanged for several minutes after. This
applies to the feed-bed behavior within the rotary kiln itself since the
rotary cylinder in itself is a relatively poor heat exchanger.

7.3 HEATING THE FEED


Radiation of heat from the hot gases to the feed bed occurs only at the
surface of the bed that is exposed to the gas. Thus the temperature of the
bed is lowest in the center and highest on the surface. In the burning zone
TI-IE ROTARY CEMENT KILN

the kiln feed, being now in a rather sticky condition, is in a constant state
of agitation as, aided by the rough and uneven kiln coating, it rises along
the upward-moving side of the kiln, then tumbles back. Because of this
tumbling action, the surface layer is constantly being folded back into the
mass of the feed, where the hot particles then transfer heat to the colder
particles by conduction. Meanwhile, new particles are being exposed to
radiation from the hot gas and the process is continuously repeated. In the
calcining zone and toward the back end of the kiln, where the surface of the
lining is much smoother and the feed still in a more or less fluid state, the
bed is turned over very little, as it slides in a zig-zag course down the kiln.
The bed is first lifted up the kiln wall, then when it reaches a certain
height it slides down and forward without turning over to any extent.
However, the turnover is larger in kilns with pelletized feed, such as wet or
semiwet process kilns.
Heat exchangers, in the form of chains, steel cylinder chambers, or
cross section, are employed in all wet-process kilns and are now being used
in dry-process kilns also. When exposed to the hot gases they rapidly
reach a high temperature and, once they come into contact with the feed
bed, transfer heat to the material by conduction..
A similar action takes place when heat transfers from the kiln wall to
the feed. The flame radiates heat to the coating adhering to the kiln. Part
of the heat then radiates to the bed, and part is transferred to the feed by
conduction when the wall turns into the bed. Fig. 7.1 shows that the wall
temperature is lowest when it emerges from the bed and highest imme-
diately before it comes into contact with the feed bed The slower the kiln
speed and the smaller the cross-sectional loading (feed) of the kiln, the
larger will be the temperature difference between these two points.
Other things being the same, there is an important relationship between
the wall effect described above and the rotational speed of the kiln, as
higher kiln speeds are more favorable than slow speeds for the heat ex-
change because of the smaller temperature difference between kiln wall and
feed bed These details must be taken into consideration by the kiln oper-
ator in deciding whether the kiln speed can be increased at any time.
In the preheater vessels this material-to-gas temperature relationship
works differently. Here the heat exchange takes place more readily and
thus the temperature difference between these two is much smaller. When
discussing heat transfer in a kiln attention must be focused upon the decar-
bonation rate of the kiln feed for this requires by far the largest amount of
heat in the entire process. In dry- and wet-process kilns, the total amount
of heat for decarbonation must be transferred to the feed within the rotary
HEAT TRANSFER

Fig. 7.1 Wall temper&me of the kiln in the burning zone, reaching a maxi-
mum just before it comes in contact with the feed bed as the kiln turns, drops
about 400 F (200 C) as it gives up much of its heat to the bed of material.

kiln itself. Since these cylinders are so inefficient in accomplishing this


heat transfer, these types of kilns must be extremely long and wide. In a
preheater kiln, the feed is calcined up to 3CL35% in the preheater tower and
thus the heat required, in terms of k&kg clinker, within the rotary kiln
proper, is much less. This makes it possible to obtain either higher pro-
duction rates on kilns of equal size or to maintain the same production rate
by sizing the kilns much smaller. With the advent of the precalciner kiln
this heat transfer efficiency has been taken a step further. On these kilns,
the kiln feed is up to 90% calcined when it enters the kiln, in other words,
most of the thermal work is done in the preheater tower. Theoretically, it
would be possible to achieve even higher calcination rates in the precal-
ciner but this is practically not possible as this would lead to the forma-
tion of viscous clinker phases and subsequent plug-ups in the lower stage
of the preheater. Most operators therefore limit their precalciner to a maxi-
mum of 90% calcination prior to the feed entering the rotary kiln to pre-
THE ROTARY CEMENT- KILN

vent the material from sticking to the walls of the preheater tower.
Preheaters and precalciners have the advantage that the heat transfer is
different from the transfer conditions found within the rotary kiln itself. In
the kiln, the material bed runs practically undisturbed along the bottom of
the kiln where the hot gases primarily heat only the surface layer of the
bed. Hence, the contact between gas and material, i.e., the heat exchange,
is relatively inefficient and takes a long time to complete. In the preheater
and precalciner, the material is in suspension resulting in an immediate
contact of feed particles with the hot gases.
This can be explained in simple terms when one looks at the time
factor that is associated with calcination in these various kilns. A dry-
process kiln requires approximately 45 m (150 ft) to achieve 90% cal-
cination and the feed takes more than an hour to travel through this section
of the rotary kiln. On the other hand, in the precalciner kiln, the same
amount of thermal work is done in less than a minute. The feed in a pre-
calciner and preheater kiln is much more uniformly calcined leading to
better and more uniform kiln operating conditions than in the rotary kiln.
Despite the sophisticated appearance of the preheater and precalciner kilns
compared to the “old” wet and dry workhorses, they are easier to control
from an operator’s viewpoint. It can be safely stated that a good wet- or
dry-kiln operator will also make a good preheater operator provided that he
is properly retrained.
There have been a vast number of technical papers published by various
authors on the subject of temperatures within different kiln systems.
From these, comparative heat profiles can be drawn that approximate the
temperatures typically found in such kilns. These are shown in Figs. 7.2
through 7.6.
A glance at these heat profiles shows the great difference in coating in
relation to the overall kiln length. In the precalciner kiln, coating takes up
45-60% of the total kiln length whereas in the wet-process kiln this
coating length is only 12-20% of the kiln length. Also significant is the
temperature differential between material and gas in the preheat zones of a
wet and a preheater kiln. It clearly shows the aforementioned superior heat
exchange in a preheater cyclone when compared to the rotary cylinder.
Detailed discussions of the mathematics involved in the different types
of heat transfer are beyond the scope of this book. It would not serve a
kiln operator in the daily performance of the job. There is, however, one
area of paramount importance to the kiln manager that deals with an essen-
tial part of kiln control; it is the heat transfer that takes place through the
- -I- -
FEED PREHEATING - . - CALCINING CLINKERINC COOLING
_ 4200

- 3850

_ 3500

_ 3150
1600
_ 2800

_ 2450

- 2100

- 1750

y 800 ;;I
,D
_ 1 4 0 0 0”: 5
v 600 - 1050 n g

_ 700 R

Fig. 7.2 Dry-process kiln material and gas temperatures.


c Id l-l-
YIUWTINC CALCINING CLINKERING COOLING 4550
Zboo
4200
2200
,I 3850
\
2000 /' 1
t \ 3500
1800 \
1'
\
\ 3150
1600 \
\ is00

2450

2100 =
g
;,
1750

1400

600 1050

700

350

Fig. 7.3 Preheater kiln material and gas temperatures


-1 4550
(r------cl-
EVAPORATION FEED PREHEAT CALCINATION CLINKERING COOLING
2400
4200
2200 r--
/
/ \ 3850
\
\ 3500
/ \
1800 \ 3150
/ \
1600 _ \
/ 2800
/ \
1400 - 2450

1200 -

u 1000 _
f
.a 8 0 0 -

600 -

400 -

200 -

/
I
20 40 60 80 100 120 140 160

20 3.0 4-O 50 m
I 1 ’ l-0

Fig. 7.4 Semidry (Lepol) kiln material and gas temperatures


L4
-I- -1-h-
2400 iGil-‘-
liVAl’OKATlON l:EED PHEIIUT CALCINATION CLINKEKING CWLINC
4200
KElll%I CL\
2200 ’ \
. I \ Jb50
I \
2000 I \
3500
\
\
/ I 3150

lb00
2800

L) 14i)i) 2450 5
L
::
j 1200 2100 k E5

2
1000
1750 5

1400 i2

600 1050 5

3
700
E

350 5

76 ft
h

Fig. 7.5 Wet-process kiln material and gas temperatures


-91-p (-; -I-
PREHEATING CALCINATION CLINKERING COOLING 4550
2400
4200
2200

3850
2000
3500
1800
3150
1600
2800

“2; !j
2450

2100

z
1750
800

1400
600
1050 %I

700 E

350

I 1 8 1
20 40 60 80 100 120 140

In

Fig. 7.6 Precalciner kiln material and gas temperatures.


THE ROTARY CEMENT KILN

refractory walls of a rotary kiln which in turn governs the temperature of


the kiln shell. Kiln managers frequently have to deal with questions that
involve the selection of appropriate refractory types and sizes to minimize
heat losses through the kiln wall and to hold the kiln shell temperature
within acceptable limits.
Some of the fundamental steps and formulas that allow for the
determination of shell temperature and heat losses when different refrac-
tories are used in a rotary kiln are discussed below.

7.4 FUNDAMENTAL FORMULAS FOR


HEAT-TRANSFER CALCULATIONS

The fundamental law for heat transfer

Q = KIA ‘l-t2
X
(7-l)

only applies to conditions where K1 and A are independent of t and x.


However, as with most other materials, the heat-transfer coefficient K1 for
refractories changes with the temperature. For this reason, the value K1
must always be accompanied with a statement as to the temperature at
which the value is expressed. For practical purposes in refractory heat-
transfer calculations, the relationship K1 vs t in most instances is linear
and therefore the heat-transfer equation can be written in the form:
t1-t2
Q=KA (7-a
X
where:
Q = quantity of heat flow per unit time
K = average of heat-transfer coefficients KI over the temperature
range tl - t2
A = heat-transfer surface area
t1 = hot-face temperature of refractory wall
t2 = cold-face temperature of refractory wall
x = refractory thickness

Either English, Metric, or S.I. units can be applied in this formula. The
heat-transfer coefficient is commonly expressed in the following terms:
HEAT TRANSFER

English system of units: Metric system of units:


Btu kcal

KI. h = Btu-inJft*h’F Kc h = kcal/mh”c


ft2. “F m2. “5 (7-3)
in. m

(Note: In the metric system, K is often denoted as X)

International system of units:


kJ
-
h
K c - = kJhnh”c
m2. “c
m

To simplify heat-transfer calculations, it is customary to use the thermal


resistance value R of the refractory wall, since R is a function of both K
and x.

R= ” (74)
k
In the above formula, K and x have the same meaning as in Eq. (7-2).
Using the R value instead of K has the advantage for the engineer to
rapidly develop graphical solutions to problems in finding the heat loss, Q
and the cold-face temperature, t2 when tl, K, x, and the ambient air
temperature ate known.
It must, however, be stressed that heat-transfer calculations for rotary-
kiln linings only &liver approximate results since several assumptions and
estimates must enter into the calculations. This can be explained in more
detail by viewing the following heat-transfer diagram of a typical cement
kiln wall.
If the direction of the heat flow is followed, problems will immediately
be experienced since it is difficult to accurately state the hot-face
temperature of the lining (tl) because both the interior temperature, tk and
the coating thickness are unknown entities in practial application. This is
especially true for the burning zone but applies as well to all other zones
of the kiln. Furthermore, although refractory manufacturers supply heat-
THE ROTARY CEMENT KILN

transfer coefficients (K) on new bricks, it is known that the thermal con-
ductivity properties can undergo changes due to the interaction of kiln
gases and the kiln-feed components with the refractory hot face during its
normal service life in the kiln.

Step 1: The first calculations are made by estimating the brick


hot-face temperature and stating the desired shell tem-
perature one is striving for. From these two, the esti-
mated temperature gradient (tt - r2) is calculated. This
is a preliminary step to find the appropriate thermal-con-
ductivity factor, K.
Step 2: In order not to introduce any more errors into the calcu-
lations, it is customary to use the logarithmic mean
temperature of the brick for calculations involving lin-
ings in cylindrical vessels.

log mean temperature = 0.434 - (7-5)


log ”
t2

When detailed information for this heat gradient are not


available, Table 7.1 can be used as a general guideline.
Step 3: From the manufacturer’s data sheet the thermal-conduc-
tivity factor (K) that corresponds to the above calculated
log mean temperature is found (or refer to Fig. 7.2).
Step 4: The wall resistance factor, R, is calculated:

where
x = lining thickness.
Step 5: Read the desired results for t2 and Q directly from Fig.
7.7 or Fig. 7.8 depending on the R factor calculated
above.
The amount of heat lost by radiation and convection from the shell can
vary considerably from kiln to kiln even if the shell temperature is the
HEAT TRANSFER

same. The values obtained by using the attached graphs are applicable to
conditions where the surrounding air temperature is taken as a constant
21 C (70 F) in still air. It is also known that different colors and shell
surface characteristics can produce lsifferent emissivities for radiation from
the shell. For example, a kiln shell painted with aluminum paint shows
an emissivity of = 0.60 whereas a typical steel shell, slightly oxidize&
shows an emissivity of between 0.80-0.88. For the purpose of this discus-
sion, a value of 0.82, is used as a constant.
Convection losses, likewise, cm dso vary depending on the character-
istics of the fluid film immediately adjacent to the kiln shell. Shell cool-
ing fans, wind velocities, and temperatures of the Smoun&g ambient air
play an important role in the amount of heat lost by convection from the
shell. A roof over a kiln section, another operating kiln nearby, and a
variety of other reasons can cause modifications in the amount of heat lost
from a given kiln section. All these intangible factors reemphasize hat
results obtained by this method of calculation are to be viewed as approxi-
mate values only.
The graphs in Figs. 7.7 and 7.8 have been developed by using two
fundamental heat-transfer equations. First, for the heat transferred through
the wall by conduction:
t1 - t2
Ql = (7-7)
R

Second, for the heat lost from the kiln shell to the surrounding air by
radiation and convection:
5
6
Q2 = 0.472 E 0.5254(t1 - 12) (7-8)

Finally, in drawing the heat-transfer curves on the graph, the individual


reference points must satisfy the relationship expressed in @s. (7-7) and (7-
8). In other words:
Q = Ql = Q2
where:
QI = amount of heat transferred by conduction through the wall
(kcal/m2-h)
kit n refractory

Fig. 7.7 Temperature gradient - refractory wall.


HEAT TRANSFER

Fig. 7.8 Shell temperature and heat losses at R - 0.20 to 1.60.


THE ROTARY CEMENT KILN

Q2 = radiation and convection losses from the kiln shell


(kcaVm2.h)
‘1 = hot-face temperature (“C)
$3 = shell temperature (T)
R = thermal resistance of wall
E = emissivity = constant: 0.82
Tabs = absolute temperature.

Example 1: A rotary kiln is being equipped with a 225~mm high, semi-


insulating lightweight lining in the upper part of the calcining zone.
What is the approximate shell temperature and heat loss?

TABLE 7.2
TYPICAL HOT-FACE TEMPERATURES
AND TEMPERATURE GRADIENTS IN
CEMENT KILN WALLS (estimates)

Hot-face Shell
temp. temp. At Log-
4 t2 01 -tz) te?np.

Kiln Hood
Wet-process kiln 650 160 490 350
. Dry-process kiln 875 200 675 460
Suspension P.h. 1100 230 870 560
FYecalciner kiln 1200 245 955 600
Grate Coolers
Lower part 300 95 205 180
Upper pa 1000 210 790 510
Kiln Outlet Zone
Without coating 1300 410 890 770
Light coating 1050 365 685 650
Medium coating 650 280 370 440
Heavy coating 450 225 225 330
Burning Zone
Without coating 1475 440 1035 860
Light coating 1100 360 740 660
Medium coating 750 280 470 480
Heavy coating 500 240 260 350
HEAT TRANSFER

TABLE 7.2 (co&d.)


TYPICAL HOT-FACE TEMPERATURES
AND TEMPERATURE GRADIENTS IN
CEMENT KILN WALLS (estimates)

Transition Zone
Without coating 1300 410 890 770
Light coating 1050 365 685 650
Medium coating 650 280 370 440
Heavy coating 450 225 225 320
Caicining Zone
Lower part:
with insulation 1230 160 1070 530
without insulation 1230 240 990 610
Middle part:
with insulation 1100 150 950 480
without insulation 1100 220 880 550
Upper part:
with insulation 1000 140 860 440
without insulation 1000 210 790 510
Preheat Zone
Suspension P.h.:
with insulation 925 140 785 420
without insulation 925 200 725 470
Dry kiln:
with insulation 875 130 745 390
without insulation 875 190 685 450
Wet kiln:
with insulation 810 125 685 370
without insulation 810 185 625 420
Chain Zone
Wet-process kiln 300 95 205 180
Dry-process kiln 575 140 435 310
Feed Inlet Zme
Wet-process kiln 205 80 125 130
Dry-process kiln 425 130 295 450
Suspension P.h. 875 205 670 460

Solution:
a) From Table 7.2, tl = 1000; t2 estimated = 140 and log mean tem-
perature of the temperature gradient is 440 C.
THE ROTARY CEMENT KILN

b) From Fig. 7.8: K at 440 C = 0.36.


C) R = 0.22510.36 = 0.625
From Fig. 7.9: Starting vertically from hot-face temperature 1000
C, find the location where this straight vertical line intercepts the
heat transfer curve near R = 0.60. Horizontal to the left, the shell
temperature is read as 120’ and horizontal to the right of this
intercept the heat loss is read as 1450 k&m* - h.

Example 2: What would the heat losses and shell temperatures be if in


Example 1 a 40% fireclay brick of the same thickness is used?
Solution:
a) tr = 1000, f2 = 210, log mean temperature, 510 C
b) K at 510 C = 0.94
cl R = 0.22510.97 = 0.24
d) From Fig. 7.3, vertical from 1000 C and horizontal from intercept
near R = 0.25, a shell temperature of 213’ and a heat loss of 3270
k&m* * h are obtained.

Example 3: An engineer desires to use a basic magnesite-chrome brick


over the second tire section which is located at the transition zone. What
height of lining is required if the shell shall not exceed 370 C? It is
projected that this area will contain a light coating.
Solution:
a) From Table 7.2, tl = 1050 C.
b) From Fig. 7.9, finding intercept tl 1050 and r2 370: R = 0.075
1050 - 320
c ) Log mean temperature = 0.434 = 613 C
1050
log -
320
d) From Fig. 7.2, K for 80% MgO brick at log mean temperature of
613 C = 2.7
e) From Eq. (74), R = x/K, determine X:
x = 2.7 X 0.075 = 0.203 = 200 mm
Thus, to maintain the temperature below 370 C, the lining thickness
should be 200 mm.
HEAT TRANSFER

‘, _Ij,‘, !
:

"ot pace temperature (deg.CI

Fig. 7.9 Shell temperature and heat losses at R - 0.02 to 0.25.


Temperature (deg. C)

Fig. 7.10 Coefficients of thermal conductivity &Urn. h . “C).


HEAT TRANSFER

Example 4: A kiln shell usually shows a faint dark red color at night
when its temperature registers approximately 460 C. Referring to
Example 2, at which brick thickness can such a red spot be expected with
40% alumina fireclay brick?
Solution:
a) From Table 7.2, tl = 1000
b) From Fig. 7.9, finding intercept tl 1000 and t2 4601 R = 0.038
1000-460
C) L.og mean temperature = 0.434 =694C
log z
d) From Fig. 7.7: K for fbeclay brick at log mean temperature of 694
c = 0.99
e) From Eq. (7-4): R = XlK
n = 0.99 X 0.038 = 0.038 = 38 mm

Thus, the shell could show a faint dark red color when the brick thickness
has worn down to approximately 40 mm.

References

1. De Keyser, W. L., 1955. Reactivity in the solid state between the


oxydes of the cement system.
2. Margin, G. 1932. Chemical eng. and thermodynamics applied to
the cement rotary kiln.
8.

Heat Balances

Before modern developments in kiln technology, there were efficient


kiln operators that never had a complete understanding of what actually
took place within a given kiln system These operators were proficient in
controlling the temperatures, pressures, and material flow rates and did
their job so well that some of them remained kiln burners for the rest of
their career. Still today, there is no absolute requirement for an operator to
be knowledgeable about heat balances. These are usually of more interest
to the plant engineer.
Nevertheless, it is appropriate to devote a separate chapter to this topic
in order to gain a deeper understanding of the process being controlled.
The operator who can keep calm and adjust to unusual situations without
losing composure deserves credit. For most erratic actions there is a
plausible explanation and these explanations can become legitimate
reasons when a deeper understanding of what really goes on inside a kiln is
reached.
Kiln operating efficiency covers a broad spectrum of considerations,
each one having equal importance. On one occasion interest in the output
efficiency may be of concern, at another time concern might shift to
cooling efficiency because the clinker cooler might not be doing the kind
of work it was designed to do. On an almost daily basis consideration
must be given to how efficiently the fuel, given to the kiln to do the true
thermal work within the system, is being used This chapter focuses on
the fuel factor because it is one of the major cost factors in the manu-
facturing process of cement. It might surprise some kiln operators when
they learn that the fuel bill for the kiln alone can amount to 3-4 million
HEAT BALANCES

dollars in a year. Not so long ago, fuel prices were so cheap in the United
States that nobody was concerned with the quantity of fuel burning in the
kiln. Today, this aspect can not be ignored
A heat balance, simply stated, consists of compiling all the heat that is
given to the kiln and then comparing this total to the total of thermal
work done and heat losses that occur in the system Whatever heat put
into the kiln (INPUT) must be accounted for in one way or another by the
heat that goes out of the system (OUTPUT). To do this requires actual
testing of the system under normal operating conditions. Some plants
have done this by means of very elaborate and sophisticated instruments,
others have used average operating data from the kiln operator’s log to
compile and calculate heat balances.
The following shows three different heat balance models, one each for a
wet, a dry, and a preheater kiln. In comparing these the differences in over-
all specific heat consumption, heat requirements, and heat losses between
these kilns can be observed.
For the uninitiated, it should be noted that the actual heat required for
clinkering the feed, i.e., the true thermal work required, is only about 431
kcal/kg (1.55 MBtu/sh.ton). In the kiln models presented, the wet-process
kiln with its 5.022 MBtu/sh.ton (1396 k&/kg) specific heat consumption
is thus only about 31% efficient, the dry-process kiln at 4.295 MBtu/-
sh.ton (1194 k&kg) is slightly better at 37%. The best fuel efficiency at
50% efficiency is found in the preheater kiln with 3.106 MBtu/sh.ton (863
big).
The best reported efficiencies are near the 2.7 MBtu/sh.ton (750
k&/kg) levels all of which are found in either suspension preheater or
precalciner kilns. This would represent a 57% efficiency which by itself
approaches close to the optimum possible limit since the most efficient
heat-transfer engine in the world works with an efticiency of around 68%.
In short, there are limits to how efficient a kiln system can become.
When kilns approach these limits it becomes increasingly more difficult
and almost impossible to squeeze an extra kJ, kcal, or Btu out of the
system to perform useful thermal work. Expressed in another manner, it
most probably would be easier to lower the fuel consumption on a wet-
process kiln from, e.g., 6.8 MB&t/t to 5.9 MBtu/t than to squeeze another
50,000 Btu/t out of it when the kiln is aheady operating at 4.95 Btu.
It should be of interest to the reader to compare the right column, which
expresses the heat output in percentages, for each of these three kiln
models. Each one of these given percentages are revealing by themselves.
But there is more to it than just testing the kiln to answer the question:
THE ROTARY CEMENT KILN

“Where does all that heat go?” Heat balances are the foundation for an
engineer to identify the opportunities for possible heat savings. There are
several areas where heat could possibly be saved to make a kiln operate
more efficiently in matters of energy consumption. Some of these
potentials are:

1. Lower kiln exit gas losses by:


installing more chains and inserts such as trefoils, lifters, and
slurry preheaters on wet- and dry-process kilns.
reducing the clinkering zone and extending the length of the
calcining zone.
operating the kiln at optimum oxygen levels with minimum
excess air in the kiln exit gases.
designing proper fuel burners that deliver optimum flame shapes
and temperatures.
on dry kilns: using the high exit gases to produce steam and
generate electrical power or using this excess heat for drying raw
materidS.
on preheater kilns: adding additional preheater cyclones or
improving the heat exchange between gas and solids within
existing cyclones.
operating preheaters with only as much bypass as is absolutely
necessary to control the alkali in the clinker.
equipping the dry-process kiln with a single-stage preheater.

2. Lower the heat losses associated with evaporating the free


water from the kiln feed in wet-process kilns by:
l using slurrythinners.

l grinding the slurry consistently to the lowest possible moisture

content.
l using filter presses to &water the slurry before it enters the kiln.

3. Lower the dust in the exit gases by:


l improving the chain system and other insert &signs in dry- and

wet-process kilns to prevent escape of dust at the feed end.


l optimizing kiln operating procedures and kiln loading to
minimize turbulence and excessive gas velocities within the kiln
itself.
TABLE 8.1
HEAT BALANCE FOR THE WET-PROCESS KILN

1ooo’S 1000%
Heat input B&/t Percent Heat output Btu/t

Combustion of fuel 4845.84 96.5 Theoretical heat required 1534.16 30.5


Sensible heat in fuel 4.25 0.1 Exit gas losses 646.45 12.9
Organic matter in feed - - Evaporation of moisture 1925.64 38.3
Sensible heat in feed 97.86 1.9 Dust in exit gas 9.73 0.2
Sensible heat in cooler air 65.17 1.3 Clinker discharge 48.68 1.0
Sensible heat in primary air 7.99 0.1 Cooler stack losses 163.29 3.3
Sensible heat in infiltrated air 0.00 0.1 Kiln shell losses 582.71 11.6
Losses due to calcination 35.03 0.7
of wasted dust
Unaccounted losses 76.44 1.5
Total 5022.11 100.0 I Total 5022.11 100.0

Note: Unaccounted losses are calculated by difference to make the two sides equal.
TABLE 8.2
HEAT BALANCE FOR THE DRY-PROCESS KILN

Heat iqwt Btu/t Percent Heat output Btult Percent

Combustion of fuel 4144.12 96.5 Theoretical heat required 1573.19 36.6


Sensible heat in fuel 3.63 0.1 Exit gas losses
Organic matter in feed - - Evaporation of moisture 1189.61
258.10 27.7
6.0 5
Sensible heat in feed 68.43 1.6 Dust in exit gas 11.14 0.3
Sensible heat in cooler air 70.30 1.6 Clinker discharge 52.62 1.2 B
Sensible heat in primary air 7.38 0.2 Cooler stack losses 508.24 11.8 z
Sensible heat in infiltrated air 0.94 - Kiln shell losses 521.17 12.1
Losses due to calcination of 15.85 0.4
wasted dust
Unaccounted losses 164.87 3.8

Total 4294.80 100 I Total 4294.80 100

Note: Unaccounted losses are calculated by difference to make the two sides equal.
TABLE 8.3
HEAT BALANCE FOR THE SUSPENSION PREHEATER KILN
Heat input Btu/t Percent Heat output BtuIt Percent
I

Combustion of fuel 2979.00 95.9 Theoretical heat required 1541.500 49.6


Sensible heat in fuel 2.61 0.1 Exit gas losses 427.370 13.8 Ei
Organic matter in feed - - Evaporation of moisture 202.210 6.5 5
Sensible heat in feed - 52.11 1.7 Dust in exit gas 1.110 - $
Sensible heat in cnoler air 66.46 2.1 Clinker discharge 56.610 1.8 r-
Sensible heat in primary air 526 0.2 Cooler stack losses 527.800 18.4 5
Sensible heat in infiltrated air 0.51 - Kiln shell losses 150.722 4.9
5.321 0.2 E
Lames due to calcination
of wasted dust
Unaccounted losses 148.302 4.8

3105.94 100 Total 3105.94 100

Note: Unaccounted losses am calculated by difference to make the two sides equal.
THE ROTARY CEMENT KILN

l returning (reclaiming) the maximum possible amount of kiln dust


to the kiln by means of insufflation, leaching, or other means.
l enlarging the feed end and the shell in the chain section to lower
the gas velocity in these areas.
4. Lower the clinker-discharge temperature by:
0 operating the clinker cooler at the minimum necessary air-flow
rates but doing the maximum possible cooling work. (This is
primarily governed by the design of the cooler and its components
which, even today, leaves a lot to be desired and offers plenty of
room for improvement)
5. Lower the cooler stack (exit) losses by:
l (same as mentioned above)

l recycling excess cooler air from the lower cooler area to the upper

cooler compartment fans. (Watch the wear on the fan blades.)


0 reclaiming excess cooler air for drying of raw materials, coal
grinding, heating of fuel oil or for combustion air in the auxiliary
flash furnace of a precalciner.
Note: Cooler stack losses am closely related to amount of
combustion air needed in the kiln. Unfortunately, the more
efficient (in specific fuel consumption) a kiln becomes, the higher
the cooler exit losses will be. The heat-balance examples show
this clearly where the inefficient wet-process kiln shows 3.3%
cooler losses compared to the more efficient preheater kiln which
shows 18.4% losses.
6. Lower kiln radiation losses by:
. using insulating refractories in the areas above the burning zone.
(Watch the maximum permissible service temperature of these
bricks.)
. selecting the proper mix design and refractory that delivers heavy
coating in the burning zone without leading to excessively large
ring formations.
. using back-up insulation on all equipment linings outside the
rotary kiln, e.g., the air ducts, kiln hoods, cooler walls, and the
preheater tower.

7. Lower cold-air inleakage into the kiln system by:


l closing up all unnecessary openings in the kiln hood, cooler, and

kiln feed end.


HEAT BALANCES

l providing for efficient kiln seals both at the lower as well as the
upper end of the kiln.
l operating the kiln with as high a primary air temperature as is
safe and possible.
Most of the above-mentioned improvement factors are part of the
inherent designs of the kiln systems where a kiln operator has usually
little influence. It is primarily a concern to the engineering department.
There are, however, many other factors, directly under the control of the
operator, that will have a bearing on the overall kiln-energy effkiency.
These factors are discussed in more details with the discussion of the actual
control functions of the kiln.
When striving to lower the energy requirements on a kiln, it must be
remembered that energy saved in one particular area might lead to a higher
energy loss in another. For example, adding more chains on a dry or wet
kiln might help to reduce the kiln exit-gas losses but invariably raises the
kiln drive-power requirements. An engineer must therefore always balance
the obtained savings by the resultant, possibly higher energy requirements
this action might cause before proceeding with the project. From an oper-
ator’s viewpoint it might be more meaningful to explain the theoretical
savings obtainable in monetary terms.
Using a computer program, a simulation was run on a wet-process kiln
to show the effect on overall fuel efficiency when certain changes in the
process variables were made. Here then are the results of this particular
run:

Wet Kiln (129.5 x 3.43 m)


Annual clinker production: 212,800 metric tons
Fuel costs: 0.5418 $ per million kcal

Change kcaltkg A?lUll


Variable saved $ saved

Exitgastemperature 20 C lower 23.86 43,200


Moisture in sluny 2.5% lower 50.94 92,500
Clinker discharge temperature 40 C lower 6.96 12,600
Cooler stack temperature 20 c lower 12.95 23,500
Kiln-shell temperature 20 c lower 16.11 29,200

Total potential savings 110.82 201,000


THE ROTARY CEMENT KILN

This example shows the resultant savings for only one relatively small
kiln. There have been reports where up to a million dollars annually have
been saved on larger kilns after extensive modifications were ma& on exist-
ing systems. Naturally, in such plants energy conservation was a high-
priority item and a well-defined conservation program must have been in
place. The middle 1970’s was a difficult period when the so-called energy
crisis occurred. At the time of this writing (1985) fuel costs have gone
down and there appears to be an energy glut on the world markets.
Unfortunately, fuel conservation efforts seem to have taken a back seat
again in many cement plants because of the apparent easy availability of
fuels. This false sense of security should not exist for the simple reason
that world market conditions change and when a kiln has a potential for
fuel conservation it should be implemented regardless of the present avail-
ability of fuel.
9.

The Chemistry of Kiln Feed and Clinker

9.1 RAW MATERIALS

The basic ingredients for portland cement consist of limestone, sea


shells, marl, or chalk, that provide the calcareous components; clay, shale,
slate, or sand, to provide the silica and alumina; and iron ore, mill scale, or
similar material to provide the iron components. The number of raw ma-
terials required at any one plant depends upon the composition of these
materials and the types of cement being produced To effect the proper
blend, raw materials are continually sampled and analyzed, and the pro-
portions adjusted as they am blended together.
After being excavated in the quarry or mine, limestone is first passed
through the primary crusher then to the secondary crusher where it is
reduced to about 3/8 in. in size. At this point other raw materials are
blended with the limestone and the blend is conveyed to raw storage piles.
Samples, mentioned above, are obtained at this point and immediately
analyzed. In a modem plant this sampling and testing is the source of data
fed into a digital computer controlling composition of stoned and blended
raw feed
In the dry process, the material is now removed from the blending piles
and delivered to the raw grinding mills, where it is reduced in size until
about 90% passes the 200-mesh screen.
In the wet process, the raw feed is transferred from raw storage piles to
the grinding mills, which are substantially the same as the ball, tube, or
compartment mills used for dry grinding. Introduction of water into the
mill along with the feed results in the formation of a slurry. After grind-
ing, dry kiln feed or slurry is drawn from storage and fed into the rotary
kiln.
THE ROTARY CEMENT KILN

9.2 CHEMICAL AND PHYSICAL PROPERTIES

Usually, any one constituent of the blended kiln feed can be found in
more than one of the raw materials. For example, typical raw materials
might contain key oxides in the proportions shown in Table 9.1.

TABLE 9.1
TYPICAL COMPOSITION OF RAW MATERIALS
CaO SiOz A12o3 Fe24 M80 LOSS

Limestone (chalk) 52.0 5.7 0.8 0.3 0.4 40.4


High-silica limestone 33.6 36.8 1.8 0.6 0.5 26.4
Cement rock 40.0 18.0 5.0 1.5 2.0 32.0
Blast-furnace slag 35.5 33.1 9.1 0.9 16.4 2.1
Shale 3.2 53.8 18.9 7.7 2.2 13.1
sad 0.8 70.0 15.0 5.0 0.2 8.6
Clay 0.5 61.0 16.9 12.4 0.4 7.8
Iron ore - 6.7 1.4 89.7 0.4 0.2
Steel-mill scale - 2.5 1.1 89.9 - 4.0

From such typical raw materials, a plant chemist tries to obtain a kiln-
feed mix that contains a predetermined oxide amount of calcium (CaO),
silica (SiOz), alumina (AlzO$ and iron (FqO$. In some locations,
mixing of only two or thme different raw materials accomplishes this,
whereas in some other plants it might need up to four or five different
materials to achieve the same results.
In addition to these basic oxides, raw materials also contain a certain
percentage of so-called impurities which show up in the kiln-feed mix.
Mugnesiu (MgO), in some plants can amount to levels of 4.2% in the mix
which, if not properly controlled, can lead to unsound (expansion) cement.
Magnesia acts as a flux at sintering temperatures which renders the burning
slightly easier. However, a magnesia-rich kiln feed tends to “ball” easily
in the burning zone which, from an operator’s viewpoint, is considered an
undesirable property.
THE CHEMISTRY OF KILN FEED AND CLIM(ER

Clinker, made from magnesia-rich feed must be very rapidly cooled once
it has been burned, to guard against production of unsound clinker. Plants
faced with this problem usually locate the burning zone very close to the
discharge end of the kiln and have quick quench compartments at the inlet
to the cooler.
Oxides of potassium (K20) and sodium (NazO), commonly referred to
as alkalies, am impurities that not only have a deleterious effect on cement
quality but can pose considerable operating problems particularly in a pre-
heater kiln. Of the two alkalies, potassium is by far the predominant im-
purity that needs close attention from the plant chemist. During the bum-
ing process, alkalies vaporize in the lower part of the burning zone, travel
with the kiln gases to the rear of the kiln, and condense again at a gas tem-
perature of around 900 C (1650 F). These alkalies react in the colder part
of the kiln with sulfur dioxide, carbon dioxide, and chlorides that are con-
tained in the kiln gases. Thus, an internal alkali cycle is created that can
lead to troublesome buildup and ring formations in the kiln. In dry- and
wet-process kilns, condensation of alkalies occurs in the lower end or just
below the chain section whereas in preheater, semidry, and precalciner
kilns this condensation takes place in the lower stages of the preheater
tower or grate preheater. Alkalies are quite an intriguing problem for any
kiln manager. There must not be too much of them in the clinker; they
should not be recycled and allowed to accumulate in the kiln; and yet they
are found in great quantities in the raw materials. To combat these
problems, various means are employed to keep these alkalies under con-
trol. In wet- and dry-process plants, part or all of the kiln dust collected in
the baghouse or electrostatic precipitator must be wasted. Some plants are
fortunate in that they have to waste only the last section of these dust
collectors, i.e., the very fine dust particles that are richest in alkalies.
Some of these plants sell this potassium-rich kiln dust as fertilizer. Pre-
heater and precalciner kilns are equipped with alkali bypass systems at the
preheater tower to control this internal alkali cycle.
There ate also plants that could tolerate a slightly higher alkali content
in the clinker but face the trouble of large, internal alkali cycles which call
for different solutions to inhibit the vaporization of alkalies.
Sulfa (SOS) is introduced into the kiln by the raw materials and the
fuel. This impurity will also vaporize to form sulfur dioxide (Sai, at a
temperature of = 1000 C (1832 F) and condense in the form of sulfates
within the kiln system They readily combine with calcium to form cal-
cium sulfates and potassium to form potassium sulfate both of which are
THE ROTARY CEMENT KILN

the prime culprits for ring and buildup problems in the upper half of the
kiln system If there is a lack of alkalies present with which the sulfur
dioxide would combine to form alkali-sulfates then much of the sulfur
dioxide would leave the kiln system with the kiln gases.
Experience gathered by many plant operators has shown that there
should be a delicate balance between the alkali and sulfur contents in the
raw mix. If the molecular ratio of alkdies-to-sulfur is significantly below
1.0 this gives rise to calcium sulfate buildups near the kiln inlet in pre-
heater kilns. By raising this ratio to 1.0 (in the form of adding alkali-rich
raw material) some plants have been successful in reducing the frequency
of calcium sulfate buildups in the kiln. Likewise, the converse has been
experienced, i.e., when this ratio exceeded 1.0 by a large margin. In such
cases, due to the excess of alkalies, alkali sulfate buildups occurred. In
such cases, the solution would be to lower the alkali cycle within the kiln
by wasting kiln dust or letting part of the kiln exit gas bypass the pre-
heater vessels. Another solution would be to add sulfur-bearing (SQ) raw
materials to the feed to balance the excess alkalies.
Buildup problems are usually attacked by first analyzing the material of
the buildup, determining its predominant compounds (in other words
trying to find out what compound caused the buildup) and finally selecting
a solution that would reduce formation of these deleterious compounds.
Chlorides originate primarily from the raw materials and from the coal.
For proper kiln operation, plant chemists usually try to hold the total
chloride content in the raw mix below 0.02%. Chlorides, too, vaporize
and react with alkalies to form alkali chloride. Alkali chlorides tend to
remain in the internal kiln cycle for a long time and can lead to heavy
coating and ring formation in the upper part of the rotary kiln and the
lower stages of the preheater. Chlorides, even in such small quantities as
0.02% in the kihr feed can become so troublesome on some preheater kilns
that they are forced to operate with a bypass of up to 15% at the preheater
tower,
Fluoride, although volatile like the alkalies, sulfur, and chlorides, does
not participate as readily in the internal cycle as the above-mentioned com-
pounds. Most of the fluoride leaves the kiln with the dust in the exit
gases or in the clinker.
It is primarily for these impurities that the preheater kiln didn’t find
rapid acceptance when it was fmt invented 40 years ago. Buildup prob-
lems plagued these kilns to such an extent that many managers considered
the almost daily kiln shutdowns for buildup removal a problem not worth
THE CHEMISTRY OF KILN FEED AND CLINKER 119

dealing with. Today these problems have been predominantly overcome


and the preheater kiln has rightfully taken the leadership in preferred types
of kilns when new plants are constructed.
In coal-fired kilns, a plant chemist must also consider the ash from the
combustion of coal as an ingredient in the kiln-feed mix. If a kiln fires 25
tons/h of coal having an ash content of 17%, there will be 4250 kg of coal
ash added hourly to the system. Typical coal ash consists of:

SiO, A12o3 Fez03 CaO SO3 cl, &Q, Na20


47% 27% 20% 3% trace

From a plant chemist’s viewpoint, this is a potent material in matters of


clinker chemistry modification and must therefore be closely monitored and
considered in mix calculations. Problems are magnified when the kiln
switches back and forth between different types of fuels burned e.g., when
natural gas is fired part of the day and coal fired for the rest. Since natural
gas does not contain the clinker modifying ash, the plant chemist must
find a kiln-feed blend that is suitable for both types of fuel burning. It
should be suitable from a burning as well as a quality-control viewpoint.
In such instances, the resultant kiln-feed mix is a compromise at best.
But, it is far better to compromise than to design a mix for coal firing and
then switch to natural gas without making a mix adjustment. The re-
percussions of such actions can be severe because the burnability of this
kiln feed is drastically changed the moment the switch to natural gas is
made. In this example, the kiln feed would be much harder to bum be-
cause the fluxing alumina and iron components in the coal ash would
suddenly be absent, A plant chemist of a precalciner plant must also give
the same attention to the amount of ash that enters the system at the flash
furnace.
Now, it will be assumed that all these factors have been given due con-
sideration by the chemist, and that the proper mix of different raw ma-
terials to obtain a typical kiln-feed composition (Table 9.2) is found.
(Note: The chemical analysis below is used as the basis for all remaining
discussions in this chapter. This analysis is only an example for illus-
trative purposes; every plant’s kiln feed will somehow deviate from the
one shown below.)
THE ROTARY CEMENT KILN

TABLE 9.2

KILN FEED COMPOSITION

Loss free Raw basis


(%) (W

SiO, 20.7 1 13.34


A1203 4.84 3.12
Fez03 2.78 1.79
CaO 67.43 43.42
MgO 2.58 1.66
Na20 0.22 0.14
K20 0.85 0.55
SQ3 0.15 0.10
Loss on ignition 0.00 35.60

The chemist now looks at the ash that enters the clinker and calculates
what the theoretical clinker composition will be after the kiln feed has
been burned and the ash included. Assume in this example that the clinker
contains 3% coal ash:

TABLE 9.3
EFFECT OF COAL ASH ON CLINKER COMPOSITION
Percent on loss-free basis

Kiln feed ash pokrltial clinker

SiO2 20.71 47.00 2150


A1203 4.84 27.00 5.50
Fe203 2.78 20.00 3.30
CaO 67.43 3.00 65.50
MgO 2.58 0.00 2.50
so3 0.15 trace 0.15
Na20 0.22 trace 0.83
KvO 0.85 tract? -
THE CHEMISTRY OF KILN FEED AND CLINKER

This clinker composition is only a preliminary potential wherein the


actual composition of the clinker can vary due to the volatility of the
alkalies and the sulfur. The important thing to remember is that, at this
stage, there is not a material that has cementitious (hydraulic) properties.
That is where the kiln comes in for it is here that this kiln feed is trans-
formed into clinker minerals to obtain the ultimate properties of what is
known as cement. During burning, this kiln feed forms the four main
clinker compounds:

Tricalcium silicate 3Ca0 - SiO2 = c,s (alite)


Dicalcium silicate 2Ca0 * SiO2 = c,s @elite)
Tricalcium aluminate 3Ca0 . Al203 = C3A
Tetracalcium aluminoferrite 4Ca0 * Al203 * FeZa, = C&Y

This compound composition is calculated with the heIp of the Bogue


formulas, from the potential clinker analysis above (loss-free basis).

9.3 BOGUE FORMULAS FOR CLINKER


AND CEMENT CONSTITUENTS

For a cement chemist, these formulas are the most important and fre-
quently used indicators of the chemical properties of a cement or clinker.
The constituents calculated by these formulas, however, are only the po-
tential compositions when the clinker has been burned and cooled at given
conditions. Changes in cooling rate or burning temperature can modify
the true constituent composition to a considerable extent.

a) Bogue Formulas for Cement Constituents.

IfAlF =>0.64
C3S = 4.07lCaO - (7602SiO2 + 6.718A1203 + 1.43 Fe&
+ 2.852SO$
C2S = 2.867SiOz - 0.7544CsS
C3A = 2.65A1203 - 1.692F%Q
THE ROTARY CEMENT KILN

c4m = 3.043FqO3

IfAIF =<0.64

c3s = 4.071CaO - (7.6O2SiO~ + 4.479Al.203


+ 2.2859Fq03 + 2.852S03)
w = 2.867SiOz - 0.7544C3S
CJA = 0
(C4AF + QF) = 2.lAl2O3 + 1.702FqO3

b) Bogue Fonnuhs for Clinker Constituents.

When appreciable amounts of SO3 and Mn203 are present in the


clinker, the values of the chemical analysis have to be recalculated to take
into account the amount of CaO that has been combined with SO3, the
amount of free lime present and the Mn203.
The values to be used in the Bogue formulas are:

Fe3 = &@3+Mn2a3
CaO = CaO - free CaO - (CaO Combined with S03)

To fmd the amount of CaO that is combined with SO3 as CaSO4


proceed as follows:

Step 1. If &O/SO3) = c 1.176 then not all of the SO3 is combined


With K20 aS K2SO4

S@ in K20 = O.85K2O

Step 2. Calculate SO3 residue

SO3 - SO, (in K20) = SO, (remaining)

If lNa2OlZ.Q (remain.)] = < 0.774 then not all of the remaining SO3 is
combined with Na20 as Na2S04.

SQ in Na20 = 1.292Na20

Step 3. Calculate the amount of CaO that has combined with the SO3
as caso4
THE CHEMISTRY OF KILN FEED AND CLINKER

CaO (in SO3) = 0.7[S03 - SO, (in K20) - SO3 (in Na@)]

Having determined the appropriate values for the CaO and F%Q, one
can then proceed to calculating the potential clinker constituents by using
the previously given Bogue formulas. When the Bogue formulas are used
for kiln feed compositions, keep in mind that the coal ash addition, dust
losses, and alkali cycles can alter the final composition of the clinker.
Also, it is necessary to use the analysis on a “loss-free” basis in the
calculations of the constituents.
Tricalcium silicate is an important constituent as it is responsible main-
ly for early strength development of mortar and concrete. Regular pordand
cement kiln feed has usually a C$ potential of 5262%. Kiln feed with a
potential in excess of 65% is extremely difficult to burn and has a poor
coating characteristic.
Dicalcium silicate accounts for approximately 22% of the clinker.
Because a higher temperature is required to form CsS than C#, under-
burning could result in a higher content of C$3 and a lower content of
c3s.
Tricalcium aluminate is responsible for the workability of the mortar.
The higher the C3A content, the higher the plasticity (workability) of the
mortar. This explains why kiln feed for the so-called plastic cements has a
higher C3A potential than that for regular cement in which the C3A
amounts to 643% of the clinker. Concrete containing cement high in C3A
is not as resistant to attack by sulfates in soil or water exposure as is
concrete made with low CJA cement.
Tetracalcium aluminoferrite governs the color of the cement. The
higher the content of C&F in the clinker, the darker the cement. This is
undesirable, as users almost unanimously prefer a light-colored cement.
Iron has the desirable property of acting as a fluxing agent in the kiln, fa-
cilitating formation of other compounds of the cement at somewhat lower
temperature than would otherwise be possible.
It is quite obvious that it is necessary to have a continuing analysis of
the material going into the kiln, if there is going to be adequate control of
the product coming out of the other end of the kiln. It is the respon-
sibility of the plant chemist to determine the composition of these
materials and to proportion them to produce a kiln feed that ensures a
uniform, high-quality clinker, combined with good bumability. Con-
tinuously uniform composition of the kiln feed is of greatest importance
for proper operation of the kiln.
THE ROTARY CEMENT KILN

Various systems are employed to introduce the feed into the kiln de.-
pending on whether the wet or dry process is to be used These systems
all serve the same purpose: to feed the kiln at a steady and uniform rate
with as little fluctuation as possible, which means that each raw material
must be carefully metered or measured.
Using our examples above and the appropriate Bogue formulas, the
following potential clinker compound content would be obtained:

c3s c2s CJA c4m


61.8 15.0 8.73 10.04

It must be again stressed here that the clinker compounds as calculated


by the Bogue formulas are only potential in nature. Formation of these
depends on the temperature, time of exposure, and cooling rate of the
clinker in the burning zone and in actuality are quite different than the
calculated values.
Up until now, a plant chemist has laid the foundation for the possible
quality of the cement that will be produced from this clinker. But, his job
doesn’t end here. It is his duty to reconcile this clinker with the bum-
ability and coatability of this clinker. In other words, he not only has to
concern himself with making a good quality cement but must also give
due attention to the ease at which this clinker can be burned. Later in this
chapter the microstructures of cement clinkers will be discussed and it will
be shown that a plant chemist, in today’s cement technology, should and
must also concern himself with the burning-zone environments that will
ultimately affect the quality of cement.

9.4 INFLUENCE OF FEED COMPOSITION


ON BURNABILITY

The “bumability” of a kiln feed is the relative ease or difficulty with


which the feed is changed into a clinker in the kiln; that is, it is an indi-
cation of the amount of fuel required to burn the kiln feed into a clinker of
good quality. Although it is highly desirable to produce at all times the
same composition of kiln feed, this cannot readily be done. One reason is
THE CHEMISTRY OF KILN FEED AND CLINKER

that most cement plants manufacture different types of cement, such as


high-early-strength, block and sulfate-resistant; therefore, composition of
the kiln feed must change from time to time as different kinds of cement
are being manufactured. Every time the feed composition changes, bum-
ability in the kiln will also change.
A plant chemist, when calculating the kiln-feed composition, will em-
ploy certain formulas to ensure that the finished product meets the specifi-
cations of the type of portland cement to be made. Kiln-feed compositions
are identified by a multitude of factors and indexes which are also used to
express burnability. These am discussed briefly below.

Silica Ratio.

The Silica Ratio is found by dividing the silica content by the sum of
the contents of alumina and iron in the kiln-feed blend. That is,

SO,
SIR =
Al,% + Fe?03

Increasing the silica ratio produces a clinker that is more difficult to


burn; in other words, the clinker is “harder” to burn. It is mainly the
content of alumina and ferric oxide that governs the combination of
calcium and silica at lower sintering temperatures.
At this point, it is appropriate to define the terms “easy burning” and
“hard burning.” In this text, an easy-burning kiln feed is one that requires
less fuel to bum to a clinker than a hard-burning feed.

Alumina-Iron Ratio.

The Alumina-Zron Ratio is found by dividing the alumina content of the


kiln feed by the iron content. That is,

A12o3
AlF= ~
Fe203

The higher the ratio, the harder the burning. Iron has a favorable
influence on the speed of reaction between lime and silica; therefore, one
can also say: Other values remaining constant, a higher iron content leads
to easier burning. Because both the numerator and denominator in the
THE ROTARY CEMENT KILN

equation are expressions of fluxing components, however, the alumina


alone is not used to express bumability.
This ratio indicates the quantity of initial liquid phase present during
burning. It is generally accepted that an A/F ratio between 1.4-1.6 is a
desirable optimum level and most beneficial to the burning of the clinker.
The higher this ratio, the harder the clinker will be to burn.

The Lime-Saturation Factor.

This factor has been used for kiln-feed control for many years in Europe
and only recently has also found acceptance by American cement manu-
facturers. When the lime-saturation factor approaches unity, the clinker is
difficult to burn and often shows excessively high free-lime contents. A
clinker, showing a lime-saturation factor of 0.97 or higher approaches the
threshold of being “overlimed” wherein the free-lime content could remain
at high levels regardless of how much more fuel the kiln operator is feed-
ing to the kiln.
H-A/F >0.64
CaO
LSF =
2.8SiOz + 1.65Al203 + 0.35F~O3

IfA/F eO.64
Ca6
LSF=
2.8SiOz + l.lA1203 + 0.7F%03

This lime-saturation factor when viewed in context with other indicators


is an excellent indicator of what the free-lime content will be in the clinker
when it has been burned at normal temperatures.
There have been instances in the past where a foreman, upon learning
from the lab results that the free-lime content was too high, approached the
kiln burner and asked him to burn the kiln hotter to lower the free lime. It
is true that lower burning zone temperatures deliver higher and, conversely,
“hotter” kilns resulting in lower free lime in the clinker. But, this is not
the only factor. High free lime can be associatd with too low a buming-
zone temperature only when the feed mix and the fuel burned remain
unchanged, and when the lime-saturation factor of the feed is below the so-
THE CHEMISTRY OF KILN FEED AND CLINKER

called saturation point. If for whatever reason the mix should suddenly
show a lime-saturation factor of, e.g., 0.97 or higher, it would be very
difficult for an operator to lower the resultant free lime by raising the
burning-zone temperature. Such action most likely would do more harm
to the coating and refractory than it would do any good to the clinker
quality.
The opposite has also been observed where complaints were voiced to
the kiln operator about burning the kiln too hot. The reason given was
that the free lime was consistently too low in the clinker. Again, a badly
underlimed mix, having a lime saturation of less than 0.88, tends to
deliver clinker that is low on free lime.
The point to remember is that when the free lime in the clinker is not
up to standards, a check with the laboratory should be made first to see if
the mix (lime-saturation factor) has changed. If this factor is still within
normal ranges, then and only then, is there an indication that the kiln
operator might not have burned the clinker at the proper temperature.
The permissible range of variation for free-lime contents varies among
different plants but the majority of the plants attempts to obtain values
that are between 0.4 and 1.2%. Experience has shown that when the
clinker is burned as close as possible to the 0.8% level of free lime, the
mix is then within acceptable levels.

The Hydraulic Ratio.

The hydrmdic ratio, developed by W. Michaelis over 100 years ago, is


very seldom used any more in modem cement technology for kiln-feed
control but is here included for plants that still regard this ratio as
significant:
CaO
HR=
SiO, + Al203 + FqO3

Percent Liquid.

Clinker, when burned at 1450 C (2642 F) will be in a so-called


semiliquid state. This viscous appearance of the clinker bed is a very
important control factor for a kiln operator when viewing the burning
zone. This will be discussed in greater detail later on. The percent liquid
is calculated by the Lea and Parker formula as follows:
THE ROTARY CEMENT KILN

percent liquid @ 1450 C = S.OAl& + 2.25Fe$& + MgO + alkalies

or by the formula which gives the same results:

percent liquid @ 1450 C = 1.13CaA + 1.35CdAF + MgO + alkalies

In both these formulas, the restriction applies that the MgO content is
limited to a maximum of 2%. In other words, a value of not more than
2% MgO can be used in these formulas.
Most portland cement clinkers show a liquid content of 25-27.5%.
Higher liquids produce stickier burning-zone clinker-bed appearances.
Since the percent liquid as calculated by the above formulas applies to a
temperature of 1450 C, higher temperatures give higher liquid and, con-
versely, lower temperatures result in lower contents of liquid in the clin-
ker. Also, since alumina, iron, magnesia, and alkalies are fluxes, higher
liquid contents make a clinker easier to bum.

Burnabilify Index.

Kuehl’s burnability in&x is based on the potential clinker compounds


C3S, C&F, and C3A. The higher the content of C3S with corresponding
lower contents in C&F or C3A, the harder the clinker is to burn.

c3s
Burnability Index =
C4AF + C3A

The Burnability Factor.

This factor was first introduced by this author in the original Rotary
Cement Kiln book It was brought to the attention of the author that
several chemists and engineers have made further research on the applica-
tion potential of combining the lime-saturation factor with the silica ratio
to express bumability. As mentioned in the original writeup, this formula
was developed based on pure empirical notions and observations and,
hence, was suspect in its fundamental reasoning.

Burnability Factor = lOO(LSF) + lO(S/R) - 3(MgO + alkalies)


THE CHEMISTRY OF KILN FEED AND CLINKER

The laboratories of F.L. Smidth, Copenhagen, have recently presented


their results of an investigation on this subject. Their findings are sig-
nificant since they used scientific methods to arrive at a direct indicator of
percent free lime in the clinker when the clinker is burned to 1500 C. The
effect of alkalies and magnesia on bumability have been assumed constant
in their formula, but due consideration was given to the effect of kiln-feed
fineness on bumability. The FLS formula is:

% free Iime @ 1500 C = [(0.33LsF) + (1.8S/R)] - 34.98 + 0.5~ = 0.136

where:
a = % retained on 4$t sieve (after acidification)
b = % retained on 125~ sieve (after acidification)

Other chemists have used the Peruy bumability factor for mix control
but have determined that the coefficients in the formula had to be changed
In other words:

Burnability Factor = x&SF) + y(S/R) - z(Mg0 + alkalies)

where n, y, and z were determined by multiple regression analysis.


There have been others that have pursued this question of expressing
burnability by using a meaningful index or factor.

Analysis of Burnability.

In addition to the kiln-feed composition discussed above, the operator of


a wet-process kiln has to consider also the moisture content of the kiln
feed, as this indirectly affects bumability of the clinker. With unchanged
composition of the kiln feed, a higher moisture content results in easier
burning. The reason for the change in bumability is the simple fact that
less feed is available to be burned when the moisture content is higher.
Because changes in kiln-feed composition have a large influence on kiln
operation, it is important that the kiln operator be advised by the labora-
tory well in advance of any upcoming change in composition. Another
good procedure is to note the chemical characteristics of the kiln feed every
day in the kiln log.
So far only the influence of chemical properties on bumability in the
kiln feed has been discussed. Plant operators must also pay attention to
THE ROTARY CEMENT KILN

the kiln-feed fineness as these physical properties of the feed can influence
bumability and the stability of the kiln operation. Since each plant pro-
duces clinker by using different raw materials and various types of kilns are
in use, there are no clear cut standards in matters of kiln-feed fineness that
would apply to all kilns. There is, however, a consensus among operators
that:

l The coarse fractions in the kiln feed tend to be more significant


than the finer fractions in their relationship to bumability.
Hence, each plant needs to test for and specify the maximum
limits of allowable fractions retained on the 30- or 50-mesh sieve
(300,500p sizes respectively).
l The kiln feed has to be ground consistently uniform and with as
little variations in particle-size distribution as possible on a daily
basis.

It is essential to make a clear distinction between a feed blend that tends


to give a better visibility in the burning zone and a blend that requires less
fuel to burn. Certain feed compositions create “dirty” conditions in the
burning zone. Here a kiln operator could come to the wrong conclusion
that because of the poor visibility the burning zone has cooled down. The
action of raising the fuel rate in or&r to clear the burning zone could result
in an overburned clinker.
On the other hand, another kiln-feed blend could possibly improve
visibility in the burning zone. In such a case, a kiln operator could
wrongly reduce the fuel rate because he concluded that the burning zone
was warming up. The final result of his action could then be an under-
burned clinker.
In Tables 9.4-9.7 the original potential clinker composition and the
various ratios and factors are computed. The percent free lime at 1500 C is
calculated using the F.L. Smidth formula but assuming that the kiln-feed
fineness (45 and 125~ respectively) remains constant at 0.04 and 0.11
percentrespectively.
For the purpose of illustration, changes of identical magnitude have
been made in all four main oxides to show the reader how changes affect
these various factors.
Several of the computed compositions are completely unacceptable in
terms of bumability and clinker quality but they serve the purpose of fa-
miliarizing the reader with some of the intricate aspects of quality control.
TABLE 9.4
CLINKER PROPERTY CHANGES (when basic oxides are varied by 0.25%)

SQ 21.50 21.25 21.25 21.25 21.75 21.75 21.75 21.50 21.50 21.50 21.50 21.50 21.50
Al,4 5.40 5.65 5.40 5.40 5.15 5.40 5.40 5.15 5.15 5.65 5.65 5.40 5.40
22 65.50 3.30 65.50 3.30 3.55 3.30 3.30 3.05 3.30 3.55 3.30 3.05 3.30 3.55 3.05
65.50 65.15 65.50 65.50 65.25 65.50 65.15 65.50 65.25 65.25 65.15
Mgo 250 250 250 2.50 250 250 250 250 2.50 2.50 250 250 250
AlkaliasNa,O 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83
sq 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15
Igoition loss 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40

loTAL 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58

: 61.78 15.03 6200 14.15 63.33 13.15 64.70 12.11 61.56 15.91 60.24 16.91 58.87 17.95 63.11 14.03 6413.ti
.48 60.46 16.03 59.09 17.07 60.41 16.07 63.16 13.99
(3 8.73 9.39 8.30 8.13 8.06 9.15 8.73 1.64 8.06 9.81 9.39 8.30 9.15
C,AF 10.04 10.04 10.80 10.04 10.04 9.28 10.04 10.80 10.04 9.28 10.04 10.80 9.28

silica ratio 247 237 2.37 2.44 257 257 2.50 247 254 247 240 240 254
Lime sahwation 93.22 93.60 94.04 94.52 92.84 9241 91.95 93.65 94.13 92.79 9232 9275 93.69
AIF xatio 1.64 1.71 1.52 1.64 1.56 1.n 1.45 1.56 1.85 1.71 1.52 1.77
Liquid (%) 26.25 27.00 26.80 26.25 25.50 25.70 2z 26.05 25.50 26.45 27.00 26.80 25.70
Hydraulic ratio 217 217 217 2.20 2.17 217 214 2.17 2.20 217 214 214 220
Lea blmbiity 3.29 3.19 3.31 3.45 3.40 3.27 3.14 3.42 3.56 3.17 3.04 3.16 3.43
ill&X
Perpy bumabiity 107.90 107.40 107.80 109.00 108.60 108.20 107.00 108.40 109.60 107.50 106.40 106.80 109.10
factor
PLs%flwlime 0.26 0.22 0.36 0.64 0.32 0.18 0.00 0.41 0.70 0.12 0.00 0.00 0.55

NOTE mS % tim time c&&ted on anwption that m-feat iin- f?ts,r remaim conrtpt at 0.0343.
TABLE 9.5
CLINKER PROPERTY CHANGES (when basic oxides are varied by 0.50%)
21.50 21.00 21.00 21.00 22.00 2200 22.w 21.50 21.50 21.50 21.50 21.50 21.50
5.40 5.90 5.40 5.40 4.90 5.40 5.40 4.90 4.90 5.90 5.90 5.40 5.40
z? 65.50 3.30 65.50 3.30 65.50 3.80 66.00 3.30 65.50 3.30 6 5 . 25 80 0 65.00 3.30 65.50 3.80 66.00 3.30 6 5 . 25 80 0 65.00 3.30 65.00 3.80 6 6 . 20 80 0
MgO 250 250 2.50 250 250 250 250 2.50 250 250 2.50 250 250
Alkali as N%O 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83
so, 0.15 03.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15

Igllitioll loss 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 5
TOTAL 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 %
g 61.78 15.03 6223 13.26 64.87 11.27 67.62 9.19 61.34 16.80 58.70 18.79 55.95 20.87 64.43 13.04 67.18 10.96 59.14 17.03 56.39 19.10 59.03 17.11 6 4 . 15 23 % ;

G-4 8.73 10.05 7.88 8.73 7.40 9.57 8.73 6.56 7.40 10.90 10.05 7.88 9.57 !z
Cd- 10.04 10.04 11.56 10.04 10.04 8.52 10.04 11.56 10.04 8.52 10.04 11.56 8.52 m
sicaratio 3
247 228 2.28 2.41 2.68 268 2.53 247 262 247 2.34 234 262 ff
L i i satluatioll 93.22 93.99 94.87 95.84 92.46 91.62 90.70 94.09 95.05 92.36 91.43 9228 94.16
AlFti 1.64 1.79 1.42 1.64 1.48 1.93 1.64 1.29 1.48 211 1.79 1.42 1.93 G
Liquid (%) 26.25 27.74 n.35 26.25 24.75 25.15 26.25 25.85 24.75 26.65 27.74 27.35 25.15
Hydraulic ratio 217 217 2.17 2.22 217 217 2.12 2.17 2.22 217 212 212 222
Lea bumability 3.29 3.10 3.34 3.60 3.52 3.24 2.98 3.56 3.85 3.05 2.81 3.04 3.57
index
Peray bumability 107.90 106.80 107.70 110.00 109.30 108.50 106.00 108.80 111.30 107.10 104.80 105.70 110.40
factor
Fls%freelilm 0.26 0.18 0.47 1.03 0.40 0.12 0.00 0.55 1.14 0.00 0.00 0.00 0.85

NOTE FLS % free l&m calr.uMed on awmption that kiln-fixd tinmess Factor rempint ConrtaDt *t 0.0343.
TABLE 9.6
CLINKER PROPERTY CHANGES (when basic oxides are varied by 0.75%)

SK& 21.50
5.40 20.75
6.15 20.75
5.40 20.75
5.40 22.2s 2225 22.25 21.50 21.50 21.50 21.50 21.50 21.50
h?Q
3.30 3.30 4.05 3.30 3.30
4.65 5.40 5.40 4.65 4.65 6.15 6.15 5.40 5.40
W?7
255 3.30 4.05 3.30 255 3.30 4.05 255
cao 65.50
65.50 65.50 66.25 65.50 65.50 64.75 65.50 66.25
20.83
50 20.83
50 2.50 2.50 250 65.50 64.75 64.75
250 66.25
254) 8
g
MgO
AIkaliasNa@ 0.83 0.83 2.50
0.83 250
0.83 250 2.50 250 2.50
0.15 0.15 0.15 0.83 0.83 0.83 0.83 0.83 0.83 0.83 t;
So, 0.15 0.15
0.15 0.15 0.15
0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40
0.40 0.40 0.40 0.40 0.40 0.15 0.15 0.15 0.15 0.15 z
Ignition loss
99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 si
TOTAL 99.58
61.78 15.03 6245 1238 66.41 9.39 70.54 6.28 61.12 17.68 57.15 53.03 65.75 69.88 57.82 53.69 57.66 65.91 E
;:
8.73 10.71 7.46 8.73 6.74 20.67 23.79 1204 8.93 18.02 21.14 18.14 11.92 2
CA
10.04 10.04 1232 10.04 10.04 10.00 8.73 5.47 6.74 11.98 10.71 7.46 10.00
c4m
7.76 Id.04 12.32 10.04 7.76 10.04 1232 7.76
i
silica ratio 247 220 2.20 2.39 2.80 280 256 2.47 2.70 247 2.28 228 270 >
Lime saturation 93.22
94.38 95.72 97.19 92.09 90.84 89.48 94.53 95.98 91.94
90.56 91.81 94.64 3
AIF ratio 2% 1.86 1.33 1.64 1.41 212
217 217 2.17 2.25 2.17 2125 217 28.49 27.89 24.60
Hydraulic ratio 26.25 24.00 24.60 4:; d:: $; ,“d 1.86 1.33 212 e
Liquid (96) 28.49 27.89 217 2.09 217
3.01 3.36 3.76 3.64 3.22 283 3.70 4.16 2.93 259 291 3.71
in&X
3.29 209 209 225 8
Lea burnability 0
Petay burnability 107.90 106.30 107.70 111.10 110.10 108.80
105.10 109.30 113.00 106.70 103.30 104.60 111.70
factor
FLS %freelime 0.26 0.15 0.59 1.4i 0.48 0.07 0.00 0.70 1.59 0.00 0.00 0.00 1.15

NOTE: l’IS + free lim cdculati on asumption that kiln-t&d fincmu hctor Iemaina .2ammt at 0.0343.
TABLE 9.7
CLINKER PROPERTY CHANGES (when basic oxides are varied by 1.00%)
21.50 20.50 20.50 20.50 2250 2250 22.50 21.50 21.50 21.50 21.30 21.50 21.50
5.40 6.40 5.40 5.40 4.40 5.40 5.40 4.40 4.40 6.40 6.40 5.40 5.40
3.30 3.30 4.30 3.30 3.30 230 3.30 4.30 3.30 230 3.30 4.30 230
65.50 65.50 65.50 66.50 65.50 65.50 64.50 65.50 66.50 65.50 64.50 64.50 66.50
250 250 250 250 2.50 250 2.50 250 250 250 250 250 250
0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83 0.83
0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15 0.15
0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40 0.40
t?
99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 99.58 2
50.99 56.28 67.28 ;c
61.18 6261 61.96 73.46 60.90 55.61 50.11 67.07 7257 56.50
15.03 11.50 7.51 3.36 18.56 2255 26.70 11.04 6.89 19.02 23.17 19.18 10.88 2
8.73 11.38 7.03 8.13 6.08 10.42 8.73 4.38 6.08 13.07 11.38 7.03 10.42 5
10.04 10.04 13.08 10.04 10.04 7.00 10.04 13.08 10.04 7.00 10.04 13.08 7.00
5
Silica ratio 247 211 211 236 2.92 292 259 247 279 247 89.69
2.22 91.34
222 95.12
279 E
Lime laturation 93.22 94.11 96.59 98.57 91.72 90.08 88.28 94.98 96.92 91.53
AIF ratio 1.64 1.94 1.26 1.64 1.33 235 1.64 1.02 1.33 278 1.94 1.26 235 G
Liquid (46) 26.25 29.24 28.44 26.25 23.25 24.05 26.25 25.45 23.25 27.05 29.24 28.44 24.05
Hydraulic ratio 217 217 2.17 2.28 2.17 217 207 217 2.28 217 2.07 207 228
Lea bumability 3.29 293 3.38 3.91 3.78 3.19 2.67 3.84 4.50 282 2.38 280 3.86
ill&X

Peray bumability 107.90 105.90 107.70 11210 110.90 109.30 104.10 109.70 114.80 106.m 101.90 103.50 113.M)
facta
FI.S%fieCliEiO 0.26 0.13 0.73 1.82 0.58 0.04 0.00 0.84 206 0.00 0.00 0.00 1.47

NOTE: FL3 56 free linm wk&kd on asumption that kiln-feed tinmw frctor rrmaini c=xlstmt at 0.0343.
THE CHEMISTRY OF KILN FEED AND CLINKER

Kiln-feed blends for productiotrof special cement-clinker types such as


high-early, sulfate-resistant, low-alkali, and oilwell cement will all have
somehow different properties and chemical compositions. Changing from
one type of clinker to another bum always requires special attention from
the kiln operator and advice of such a change should be given well before
this new feed is being used in the kiln.

The Liter-Weight Test.

One of the easiest tests a kiln operator can perform to learn if he has
burned the clinker at the proper temperature is the liter-weight test. Free-
lime content also gives essentially the same information but analysis for
free-lime content takes up to an hour until the results are reported. Since
the sample is usually taken from the outlet of the cooler, the results tell
what was done 1.5-2 h before which, from an operator’s viewpoint, is not
much help. In the liter-weight test, the sample is first passed through a lo-
mm screen then through a S-mm screen. The fraction retained on the 5-
mm screen is then allowed to fall through a prescribed distance into a lOOO-
ml container which is in the shape of a frustrum of a cone with the small
end up. The weight of the clinker in this container, called the liter weight,
indicates how well the clinker has been burned, as a hard-burned clinker has
a higher liter weight than a soft-burned clinker, provided there is no change
in raw-mix composition. A well-burned clinker has a liter weight between
1250-1350 g. The liter weight can vary considerably, even though the
clinker is well burned, between one feed composition and another. If the
raw-feed composition remains constant, the best clinker will have the
appropriate liter weight and lowest free-lime content. The time required to
run a liter-weight test is approximately 5 min.
Care must be taken to assure that the entire test procedure is carried out
in the same consistent manner. From personal experience it better serves
the operator to determine at what time the liter weight should be per-
formed. In a well-running kiln, there is not much use in wasting time run-
ning a liter-weight test every hour. Liter-weight tests should be done
when the operator or foreman is not quite sure if the clinker is being
burned at the proper temperature.
Here, too, as with the free-lime test, there is the disadvantage of the
time lag to consider since most samples are extracted from the cooler dis-
charge. Some plants have overcome this by extracting the sample directly
from the kiln hood before it falls into the cooler using careful cooling
procedures to guard against bums.
THE ROTARY CEMENT KILN

How many and what types of clinker indexes, factors, etc., are to be
used for effective quality control is a matter that has to be decided by the
operator and plant chemist. This author has used a backward approach to
arrive at an optimum mix that theoretically would successfully produce a
desired clinker. First, certain fmed desired properties like the lime-sat-
uration factor, silica ratio, and A/F ratio are set and using these as con-
stants the needed oxide composition is arrived at. This method of opti-
mum-mix design is very tedious when done by hand since it involves
repeated trial-and-error calculations until the right suitable mix is reached.
But, with the help of a computer program, this work is greatly simplified
and quickly accomplished. An example using this procedure is shown in
Table 9.8.
In the preceding pages the chemistry of the kiln feed and clinker have
been extensively discussed. The novice reader should now have a fairly
broad knowledge of the many problems and factors that are associated with
making cement clinker in a kiln. For new kiln operators there will come a
time when the realization that something is not quite right with the mix
occurs. It might suddenly happen while burning a problematic kiln feed.
Sometimes these problems can persist for several days and a kiln operator
can become frustrated. It is common for kiln operators, whenever the kiln
doesn’t operate andhandle properly, to blame these problems mistakenly
on the laboratory staff. In most instances this is not justified. It must be
realized that the plant chemists face just as many obstacles as the kiln
operator. Most of the time a chemist is aware that the mix is not up to
his liking but he can do nothing about it because of many factors not
directly under his control. Raw materials might not be available to make
appropriate corrections, kiln-feed reserves could be at low levels, or the raw-
grinding department might face it’s own disturbances.
In other instances the laboratory might get right back at the kiln
operator and tell him that there is nothing wrong with the mix. The best
way to guard against this type of stalemate is to establish a good dialogue
between the laboratory and the kiln control room A kiln operator can
accept problems when they are explained Being made aware of potential
upcoming disturbances in the kiln is a mom tolerable situation.
Assuming that the point has been reached where a shining clinker of
superior quality has discharged from the cooler, the stage would be set for a
good finished product. However, the cement as yet has not been produced.
If just clinker alone is ground, the final product would be an inferior,
unworkable cement. To control setting time, gypsum has to be added to
the clinker during the finish-grinding process. Typical portland cement
contains approximately 5% of gypsum.
TABLE 9.8
OPTIMUM CLINKER COMPOSITION

ENTER LOSS-FREE VALUES (typical in clinker)


Alkalies (KzO & NazO) 0.60
sulfur (SOj) 0.40
Magnesia (MgO) 2.30
Iron (F%OJ most desirable percent a =

Step 1. Set desired A/F ratio b = 1.80 < 1


Step 2. Calculate Al&X, content = a *b C = 5.67
I OK? Y N
Step 3. Set desired silica ratio d = 2.45 I +I
Step 4. Calculate SiOz content = d * (a + c) 0 21.61 OK? Y N
Step 5. Set the desired Lime-Saturation factor (Kuehl) 0.93
Step 6. Calculate CaO content
= f * [2.8e) + (1.6%) + (0.35u)] g = 66.00
I < ‘T
Step 7. Check sum total all oxides (must be > 99.5 < 100.0) 99.72 OK? Y N
Step 8. Calculate potential clinker compounds by Bogue 4-J
(when A/F > 0.64) w 60.66 OK?
CY 16.19
Cd 9.70 OK?
C,AF 9.59
Step 9. Calculate bumability factor 108.80 OK?

(Proceed with calculating coal-ash addition effect before determining the actual kiln-feed mix composition.)
THE! ROTARY CEMENT KILN

At this stage, the production of cement, the binder that holds the sand
and the aggregates in the concrete and mortar together, is completed. This
is the most durable construction material that is known. No wood, glass,
or steel construction will ever have the life span of concrete. The many
things that were made of clinker serve as tributes to the kiln operator. An
old, now retired kiln operator, told this author a few years ago that one of
the most memorable things he did in life was to be involved in making the
cement used for the construction of the Hoover Dam in Nevada.
Considered at that time to be one of the eight technical wonders of the
world, it is still around and doing well. What a monument to a kiln
operator that had no automatic controller, no computers, no air-conditioned
control room, and operated a kiln that was driven by a leather belt around
the shell.

9.5 CLINKER MICROSCOPY

It was mentioned earlier that the clinker very rarely contains the
amounts of clinker constituents as are calculated by the Bogue formulas.
Sooner or later, every plant chemist also faces the situation where a drop
in cement quality cannot be explained by the earlier mentioned clinker
indexes, the chemical composition of the clinker, or by factors related to
the finish grinding of the cement. In many respects the plant chemist can
have the same problems as the kiln operator in that once in a while things
go completely contrary to what one has learned or experienced over a long
period of time. In short, poor cement quality would seem to contradict
apparently good clinker quality and good finish-grinding parameters. Such
reasoning has led to the method of investigating the microstructures of
clinker in an attempt to identify changes in clinker quality as a result of
changes in the burning and cooling-zone conditions in the kiln. Exam-
ination by microscope can reveal things that chemical analysis of a clinker
sample will not.
Dr. Yoshio Ono of the Onada Cement Company, Japan introduced a
new technique in the early 1970’s that allowed for microscopic obser-
vations of clinkers and related these to the burning conditions in a kiln and
the prediction of the compressive strength of the finished cement. These
observations are, to this date, qualitative and not quantitative in nature. As
with any new method, the cement industry took a long time to accept it as
a valuable tool in quality control. Today there still exists some limited
THE CHEMISTRY OF KILN FEED AND CLINKER

disagreements as to the accuracy of the values found by this method in


relation to observation of the burning conditions in the kiln. What most
chemists and kiln operators, however, seem to agree upon is that micro-
scopic observations of the clinker microstructure is indeed a valuable tool
for kiln control and that only the interpretation of the results is in need of
further study and refinement. With the passage of time, when more clinker
investigations have been performed by this method, there is a good chance
that the microscope might become part of standard testing equipment for
clinker quality control. Several cement plants are using these microscopes
on a regular basis and some have even gone so far as to locate them, for
the operator’s use, in the kiln control room.
As previously discussed, C$ (alite) is the most important clinker
constituent that governs the strength of the cement. Alite forms in the
burning zone by combining in the liquid phase the C,S (be&) with
available free lime. Thus, it is important to burn the kiln in such a
manner as to obtain the maximum amount of alite and the minimum
amount of free lime in the clinker.
As this reaction proceeds, the alite crystals grow in size as the behte
crystals progressively shrink and the free lime starts to disappear. Under
certain burning zone and cooling conditions this newly formed alite can be
unstable and revert back to CzS @elite) and free lime in accordance with
the following:

C$3 + free CaO + (liquid phase) + C3S + (slow cooling) + C2S + CaO

On slow cooling, lime is extracted from the surface of the alite crystal
and leaves a belite layer (in petrographic terms this is referred to as
birefringence) around the alite.
Hard burning of the clinker as well as too long an exposure to very
high temperatures (long burning zones) leads to an increased liquid phase,
causing formation of excessively large alite crystals which in turn is detri-
mental to the quality (strength) of the cement. Conversely, insufficient
temperature leads to formation of insufficient and smaller C3S (alite)
crystals, excessive belite and high free lime which, too, leads to inferior
cement quality.
Crystal sizes are therefore important to cement quality and only micro-
scopic (petrographic) investigations of clinkers shed light on this aspect of
clinker properties.
Ono’s method calls for investigation of clinkers for the following
variables:
THE ROTARY CEMENT KILN

a) Alite size
b) Belite size
c) Birefringence
d) Belite color

The methods of sample preparation, evaluation, and interpretation of the


observations are beyond the scope of this book Special training to this
end is required. Ono has developed graphs where the observations Of the
above variables can be related to the burning and cooling conditions in the
kiln. Users of this technique report good correlations when the kiln feed
is fairly uniform but there seem to be some problems in obtaining good
and meaningful results when kiln-feed chemistry and fineness fluctuate
frequently.
By means of the microscope some plants have been able to identify the
true causes for the formation of so-called “chestnut” clinker. For years the
consensus has been that these clinker balls, having a brown center core,
were caused by reducing conditions (lack of oxygen) during the sintering
process wherein the trivalent iron (Fern) was transformed into ferrous oxide
(Fen). Work done by G.R. Long, Blue Circle Research Division (pro-
ceedings of the 4th International Conference on Cement Microscopy) has
shown that this light brownish core in a clinker can also be caused by
sintering in a normal oxidizing atmosphere with rapid cooling thereafter.
The underlying causes can be differentiated from each other by visual ob-
servation. If reducing conditions cause the brownish core, there is a clearly
noticeable yellow band which separates the brown inner core from the
outer grey-black coat. In “chestnut” clinkers that are burned nevertheless
in oxidizing environments but quickquenched, no such yellow band has
been observed.
There is no doubt that microscopic investigations of clinker have just
started to come to the forefront. Application possibilities for quality con-
trol by this method are a natural step in the continuing striving by chem-
ists and operators to improve and refine their procedures and to learn more
about the complexities of cement manufacturing.
10.

Reaction Zones in the Rotary Kiln

Controlling the kiln operation means keeping a watch on all the


different zones and reactions that take place within the kiln system The
chemical reactions and phase formations taking place within the kiln are
discussed in Chapter 7. None of these zones can be neglected or ignored if
the kiln is to be operated properly. There are five distinct zones within the
kiln, their location and length being different for each type of kiln system
used. These zones are shown in Table 10.1.

TABLE 10.1
ZONES IN ROTARY CEMENT KILNS

Temperature range of material

Drying and preheating zone S-805 60-1480


Calcining zone 805-1200 1480-2192
Upper-transition zone 1200-1400 2192-2552
Sintering zone 1400-1510 2552-2750
Cooling (lower-transition) zone 1510-1290 2750-2350

Kiln operators define the upper-transition, the sintering, and the lower-
transition zones as one complete unit, referred to as the burning zone. The
burning zone is that area in the kiln where clinker coating exists on the
refractory surface.
THE ROTARY CEMENT KILN

In the drying and preheating 20~ aJl the free water in the feed is
removed and the temperature of the feed is raised to approximately 805 C
(148 1 F) at which temperature calcination begins. During drying, the feed
temperature remains at 100 C (212 F) until all the free moisture is re-
moved. Starting at a temperature of 550 C (1022 F) the chemically com-
bined water in the clay minerals is also driven off. Thus, in this particular
zone, the reactions taking place in succession are: raising the feed tern-
perature to the evaporation temperature, evaporating the free water, libetat-
ing the chemically combined water, and raising the feed to the calcining
temperature.
From an operator’s viewpoint, the cdcining zom is the most important
zone that governs subsequent events in the burning zone. Calcination is
the action through which carbon dioxide is driven off from limestone and
magnesium carbonate, leaving free lime and magnesia:

OCo, + CaO + CO2


(limestone) (free lime) (carbon dioxide)

MsCo, + MgO +
(magnesium carbonate) (magnesia) (carboZ%oxide)

Carbon dioxide dissociates from the feed and is carried away by the kiln
gases. Kiln feed that is not completely calcined before it enters the bum-
ing zone is difficult to bum and is one of the main reasons for upset kiln
conditions. Thus, complete calcination of the kiln feed before it enters the
burning zone is essential to proper burning of the clinker.
From the above reaction of limestone it follows that the formation of
free lime in the calcining zone is accompanied by a corresponding reduc-
tion in the CaCO3 (calcium carbonate) content in the feed If samples
were extracted in incremental distances along the length of the calcining
zone, the rate of reaction taking place could be determined with reasonable
accuracy as the feed travels down the kiln. Stopping a kiln and extracting
these samples from the feed bed, however, would not serve this purpose
because the feed bed would continue to be calcined after the shutdown until
the temperature of the feed fell below 805 C (148 1 F). Research work has
been done along this line by installing sample ports on the kiln shell and
sampling the feed bed while the kiln remains in operation. The results
obtained are revealing and of interest to the cement industry but such pro-
cedures are prohibitive, because of the high costs for the day-to-day control
of the kiln operation.
REACTION ZONES IN THE ROTARY KILN

On preheater and precalciner kilns this method of checking the calcining


rate is not a problem because it is easy to sample the material as it leaves
individual cyclone stages. Most dry-process kilns also have a sample port
at the discharge end of the chain section that allows for analysis of the
degree of calcination. In this case, however, the results are not as reveal-
ing because the sample point is at a location where calcination is usually
just starting. Nevertheless, checking the calcination rate on a weekly or
monthly basis for dry kilns and on a daily basis for precalciner and pre-
heater kilns is a recommended procedure.
For a continuous indication of the calcining conditions while the kiln is
in operation, the operator monitors the gas analyzer that records, among
other gas components, the CO2 content of the kiln exit gases. Because it
is an important control function, CO, control is discussed in more detail
in the chapter on kiln controls.
The term upper-trunsition zone is relatively new to kiln nomenclature.
In the chapter on heat transfer, mention is made of the interim-phase for-
mations once the feed has been calcined. There is an area in the kiln where
calcination and interim-phase formations overlap each other. Some of the
finer particles, or the ones that were most favorably exposed to the hot
kiln gases on the feed-bed surface, can reach complete calcination and start
to react with the alumina, silica, or iron to form these interim compounds.
With other particles, calcination would not be finished and thus no liquid
formation would begin. Engineers generally refer to this area as the upper-
transition zone. This is identified as the upper third of the area where clin-
ker coating is formed in the kiln. This area is located inside the kiln,
directly behind the flame where the feed bed has a “dark” color. In the
temperature diagrams shown in the chapter on heat transfer, this area is
identified by the sudden rapid rise in material temperature just at the start
of the burning zone.
The sintering zone is where the final stages of clinker compound
formation takes place. Unstable interim phases such as C,A, and C,F
transform into the stable compounds CsA and C&F. Also, C2S com-
bines with more free lime to form the all-important compound C$.
sintering in the heart of the burning zone is an exothermic process, i.e.,
heat is evolved instead of being absorbed during this process. Sintering
takes place in the middle part of the burning zone directly under the
illuminated part of the flame. It is here where the most viscous behavior
of the clinker bed can be observed and where the material temperatures are
the highest, i.e., > 1400 C (2552 F).
The cooling (or lower-trunsifion) zone entails, usually with few
THE ROTARY CEMENT KILN

exceptions, the last 3-6 m (10-20 ft) of the discharge end of the kiln.
Contrary to popular misconceptions, the process of clinker formation is
not complete when the material has passed the hottest area. The manner in
which the clinker is cooled greatly affects its quality. Depending on the
location of the flame, cooling, i.e., solidification of the partly liquid com-
pounds, can be either slow or rapid. It is generally true that rapid cooling
is beneficial to both the quality of the clinker and its grindability. For
rapid cooling, the sintering zone is very close to the discharge end; thus
the clinker discharges from the kiln at a temperature of about 1370 C
(2498 F). As soon as it enters the first quench compartment in the cooler
it is rapidly cooled. Conversely, when the burning zone is located too far
uphill, the clinker will reside for a longer time in the lower-transition
zone, being slowly cooled, and thus discharged from the kiln at a tem-
perature of around 1200 C (2192 F).
It is now of interest to discuss how the length of the above-mentioned
zones can vary from one kiln system to another.

10.1 LONG WET-PROCESS KILNS

In a wet-process kiln, the feed enters the kiln in the form of a slurry,
having a moisture content of approximately 30%. Various types of heat
exchangers are used to remove this moisture from the slurry, the one most
widely used being chains at the back end of the kiln, shown in Fig. 2.2.
Other kiln systems dry the slurry and pelletize the kiln feed before it enters
the kiln, using the kiln exit gases for this purpose. Modem, large rotary
kilns may use a series of heat exchangers; for example, steel cylinder
drying chambers and chains. In some kilns the gases, before they enter the
chain section, pass through a section where the interior of the kiln is
divided lengthwise into four compartments over a length of approximately
20 ft. This divides the feed bed and thus provides a better exchange of heat
between the hot gases and the relatively cold feed. These devices are better
known by the term trefoils.
The slurry, when it enters the kiln, has a temperature of approximately
38 C (100 F), rising to a chain discharge temperature of between 200-
260 C (392-500 F) in kilns with chain systems that dry all the moisture
within the chains. On nodule kilns, i.e., when the feed still contains ap-
REACTION ZONES IN THE ROTARY KILN

proximately 8-15% moisture when it leaves the chain system, the feed
temperature will be around 100 C (212 F) at this location. Wet-process
kilns in which all the free moisture is driven off in the chain system are
usually more energy efficient but often show higher kiln exit dust losses,
have a tendency toward mud-ring formations in the chains, and produce feed-
bed characteristics below the chain zone that are identical to a dry-process
kiln. Kilns with chain systems that are designed to form and discharge
nodules tend to be easier to control and have less tendency to generate dust
flushes (kiln upsets) into the burning zone. From an operator’s ease-of-
control viewpoint the latter is clearly more desirable. In a typical wet-
process kiln, the various zones generally take up the following percentage
of total kiln length:

approximate % of kiln length

Slurry preheat 5%
Evaporation and preheating 40%
Calcination 33%
Burning zone (upper, sintering, 22%
and lower-transition zone)

10.2 THE DRY-PROCESS KILN

In the dry process, the feed enters the kiln in dry powder form. As in a
wet kiln, the feed passes through a chain section whose primary function
is to preheat the feed to a chain exit temperature of around 730 C (1346 F).
Thus, whatever free moisture there is in both the feed and the chemically
combined water, is driven off within the chain system Calcination starts
a few meters below the chain section. The typical length of various zones
in a dry-process kiln are:

approximate % of total kiln length

Feed drying and preheating 30%


Calcining 47%
Burning zone (upper, sintering, and 23%
lower-transition zone)
THE ROTARY CEMENT KILN

10.3 THE SEMIDRY (LEPOL) KILN

The drying, the dissociation of the chemical water, the preheating to


calcining temperature, and even partial calcination in this system takes
place outside the rotary kiln, i.e, in the grate preheater. Hence, the rotary
kiln proper has only to complete the calcination and form the clinker com-
pounds. The length of the zones in the Lepol kiln are thus quite different
from the aforementioned kilns.

approximate 940 of total kiln length

Calcining zone 55%


Burning zone (upper, sintering, and 45%
lower-transition zone)

10.4 THE PREHEATER AND PRECALCINER


KILNS

These kilns are fed dry powder feed at the upper stage of the preheater
cyclone tower. Preheating, dissociation of the chemically combined water,
and partial calcination of the feed, here too, takes place outside the rotary
kiln. In a preheater kiln, the feed enters the rotary kiln approximately
3040% calcined whereas in the precalciner kiln this feed can be up to 90%
calcined. For this reason, the burning zone in a precalciner kiln is longer
and the calcining zone shorter than on a preheater kiln as shown in the
following:

approximate 70 of total kiln length


-
Preheater PrecdCiner

Calcining zone 35%


Burning zone 40% 65%
11.

Coating and Ring Formations


In A Rotary Kiln

11.1 COATING FORMATION IN


THE BURNING ZONE

A good protective coating on the refractory in the burning zone serves


to prolong the life of the refractory. Frequent replacement of refractory
costs a large amount of money, not only because of the high cost of the
refractory, but also because of the loss in production while the kiln is
down for lining replacement. Needless to say, frequent renewal of refrac-
tory is an undesirable condition in any kiln.
Although refractories in the burning zone have to be replaced from time
to time, a kiln operator nevertheless has the capacity to increase (and
unfortunately also to decrease) the life of the lining by his ability to
control the coating in the burning zone.

Nature of the Coating.

Coating is a mass of clinker or dust particles that adheres to the wall of


the kiln, having changed from a liquid or semiliquid to a solidified state.
The solidified particles adhere to the surface of the coating (CS in Fig.
ll.l), (or the refractory surface (BS) when no coating exists), as long as
the temperature of the surface of the coating is below the solidifying tem-
perature of the particles. Coating continues to form until its surface
reaches this solidifying temperature. When the kiln operates under such
condition at equilibrium, the coating will maintain itself. This means that
THE ROTARY CEMENT KILN

theoretically no new coating is formed. When this temperature is ex-


ceeded, however, the particles on the surface of the coating change again
from a solid to a liquid state, and the coating will start to come off.

DIRECTION I OF
rn AUCI -BARREL
REFRACTORY L I N I N G
‘ING

Fig. 11.1 Heat, passing through the shell of the kiln, must be constantly
replenished by the flame in order to maintain a condition of equilibrium
necessary for coating formation.

There is a temperature drop between the coating surface (CS) and the
kiln shell (KS), the heat flowing in the direction indicated by the arrow in
Fig. 11.1. (Heat always travels from a place or body of high temperature
to a place of body of lower temperature.) This heat transfer is governed to
a great extent by the conductivity of the refractory and the coating. The
better the conductivity of the refractory, the better the chance of coating
formation, explained by the fact that the more heat that travels in the direc-
tion of the arrow, the lower the temperature will be at the surface of the
coating. Because the coating consists of particles that have changed from a
liquid to a solid state, the amount that any kiln feed liquefies at clinkering
COATING AND RING FORMATIONS IN A ROTARY KILN

temperature plays a very important role in coating formation. A kiln feed


with a high liquid content at clinkering temperatures is more effect.iVe in
coating formation than a feed low in liquid. Kiln feeds with a high liquid
phase (easy-burning mixes) have a high content of fluxes: the iron, alu-
mina, magnesia, and, alkalies. On the other hand, hard-burning mixes (low
in iron, alumina, magnesia, and alkalies, and high in silica and lime) do
not have a favorable influence on coating formation. Alkalies entrained in
the gas stream promote the formation of coating (unfortunately rings also)
because of their high fluxing characteristics. ‘
Because the surface temperature is probably the most important factor
in the formation of a coating, it is obvious that the flame itself has a
significant effect on coating formation because the shape of the flame
directly governs the surface temperature at any given point in the burning
zone. A flame that is too short, snappy, and wide can erode the coating
because of the great heat released over a short area with this kind of flame.
A long flame is more favorable to coating formation in the burning zone.
In Chapter 6 it was pointed out that short flames are desirable for better
control of the burning operation, but the flame should be shortened only to
the extent that it will not harm the coating.
Once all the factors have been taken care of and the foundation for a
good protective coating established, it is then up to the kiln operator to
control the coating during operation of the kiln. It is his responsibility to
form and maintain a good solid coating in the burning zone.

Operating Conditions.

Operating conditions are just as vital for coating formation as all the
other factors mentioned above. Assume that a kiln will be operated from
one extreme of temperature to the other, that is, a cold, a normal, and a
badly overheated kiln; that the same kiln-feed composition is burned in all
three examples; that the solidifying temperature is 2400 F (1315 C); and
that 24% liquid is formed at the point of investigation, under ideal oper-
ating condition.
First, consider the cold kiln (Fig. 11.2A). In this case almost no coat-
ing is formed The coating surface temperature as well as the feed tempera-
ture is too low to produce the necessary amount of liquid matter that would
promote coating formation. The condition in this example is commonly
referred to by kiln operators as the kiln being in a “hole.” This example
also supports the widely known fact that no new coating can be formed
while the kiln is cold.
THE ROTARY CEMENT KILN

In the normal kiln (Fig. 11.2B), enough liquid (24%) is present to form
a coating. Temperature of the coating when it emerges from the feed bed,
as well as when in contact with the feed, is below the solidifying temper-
ature of the feed particles. The particles will adhere to the wall and solidi-
fy, and will continue to do so as long as the surface temperature of the
coating remains below the solidifying temperature of 2400 F (1315 C).
Whenever the wall reaches this temperature no new coating will form.
The coating is in equilibrium.
In the hot kiln (Fig. 11.2C), because of the extremely high tem-
peratures of the feed and the coating, too much liquid is formed. As all
temperatures are above the solidifying temperature, the coating transforms
from a solid back to a liquid again. In such a condition, coating will come
off, and the feed because of its high liquid content will “ball up.” Needless
to say, this condition is extremely harmful to the kiln and to the
refractory.
Most basic refractories, and especially the dolomite liners, are not able
to withstand prolonged exposure to the high flame temperatures without
this protective coating. As was mentioned in the previous chapter, the
burning zone is divided into three subzones namely the upper-transition,
the sintering, and the lower-transition zones. Because of the lower liquid
content in the feed and because of the frequent temperature changes, the
upper- and lower-transition zones are areas where formation and
maintenance of coating is the most unstable. Shifting burning zone
locations produce a similar shift in the location where coating is formed;
thus, unstable coating conditions are most frequently observed in the upper
and lower end of the burning zone. This is clearly supported by the fact
that most rotary kilns experience the most frequent refractory failures in
these two critical areas. It should be noted that since the upper and lower
burning zones are also within the vicinity of the fmt and second tires,
brick failures am not only the result of variations in burning-zone
conditions but, are also often the direct result of excessive tire clearance
and shell ovality. Both the frequent falling out of coatings in these areas
and the formation of too much coating can lead to troublesome ring for-
mations.
Ring formations in the lower-transition zone (i.e., at the kiln discharge)
are referred to as nose rings. Others refer to these as ash rings when the
kiln is coal fried. Ring formations in the upper-transition zone are referred
to as clinker rings. These ring formations can in many instances be so
severe that they force operators to shut down the kiln and shoot these rings
A COLD FLAME B NORMAL FLAME

Fig. 11.2 Kiln A, being in a hole, is not hot enough for proper formation
of coating; good coating is formed in kiln B at normal operating temperatures;
the high temperature in kiln C causes too much liquid to form and can result in a
serious loss of coating.
THE ROTARY CEMENT KILN

out with an industrial gun. The Cardox system has been successfully used
for many years in Europe on several kilns to remove such rings. These
devices are affixed to the kiln shell in strategic locations and use CO,
cartridges to blast the rings while the kiln has only to be stopped for a
short interval to load and trigger the cartridges.
Much research work has been done on the probable causes of these ring
formations in the burning zone. The possible causes are many and no one
single factor has yet been found that would be the main cause for all the
rings formed. What seems to be true for one particular kiln might be
competely wrong for another kiln. This is clearly explained in the fol-
lowing example: On many coal-fired kilns, operators have found a relation-
ship between the fusion temperature of the coal ash and the frequency of
ring formation. There appears to be more ring formation when the fusion
temperature is low, i.e., when the ash contains larger amounts of fluxing
iron and alumina and less silica. However, this could not be the only
cause for such ring formations because natural gas- and oil-fired kilns,
which have no ash deposits in the burning zone, can have just as many
ring problems as the coal-fued kilns, Hence, solutions for the elimination
of rings in the burning zone are predominantly found by a process of
elimination. First, all probable causes are listed and then each suspected
cause is eliminated or changed until hopefully an answer is found. From
personal experience, the author has found the following factors to be
possible contributors to ring formations in the burning zone:

POSSIBLE CAUSES FOR RING FORMATION


IN THE BURNING ZONE

a) Coal fineness too coarse.


b) Low fusion temperature of coal ash.
c) Kiln feed high on liquid content (silica, A/F ratios as well as lime
saturation factor low).
d) Incomplete calcination of the feed as it enters the burning zone
(frequent dust flushes into and poor calcining conditions behind
the burning zone).
e) Frequent changes in chemical composition and fineness of kiln
feed,
f) Excessive dust generation in the cooler and burning zone
(including changes in dust insufflation rates on wet-process
kilns).
COATING AND RING FORMATIONS IN A ROTARY KILN

g) Kiln speed too slow and feed loading too high in normal
operation.
h) Excessive variations of flame temperature and length during
normal operation.
i) Frequent changes in secondary air temperatures.
j) Excessive frequency of kiln-operating upsets (burning zone
temperature and location varies too frequently and by too large a
range).
k) Increased volatility of, and frequent changes in, alkali and sulfur
contents in the fuel and feed.

Others have found other reasons for ring formation in the burning zone.
Thus, the list could possibly be expanded to over 30.
It is of interest that half of the cited factors can be somehow controlled
by the kiln operator and action taken to stabilize the flame and the kiln
operation that might be beneficial in lessening the frequency of ring
formation.

11.2 COATING AND RING FORMATION


UPHILL OF THE BURNING ZONE

Less frequent but nevertheless equally troublesome are the so-called feed
rings that form in the calcining zone of the rotary kiln. Wet-process kilns
often experience so-called mudring formations in the chain section too.
Finally, many preheater kilns and Lepol kilns experience ring formations
and build-up problems at the feed inlet and in the lower preheat cyclone
stage.
In investigations on this subject it has been found that the majority of
these rings and heavy coatings in dry- and wet-process kilns are associated
with one of the following factors:
a) Internal cycle of the volatile constituents from the kiln feed and
fuel (alkalies, sulfur, chlorides).
b) Kiln-feed fineness.
c ) Irregular and insufficient control (frequent fluctuations) of the feed-
end temperature and kiln draft.
d) Excessive dust generation within the rotary kiln proper.
THE ROTARY CEMENT KILN

Analysis of the materials from these rings or excessive coating buildup


invariably showed high contents of calcium sulfates, potassium chlorides,
and/or alkali sulfates. Efforts to alter the internal and external cycle of
volatile components in the gas or feed stream have in many instances
resulted in less frequent ring formations. Although the aforementioned
reasons apply to dry- and wet-process kilns, there have been many reports
by others that have found a similar relationship in preheater and Lepol
kih.
In one case of a wet-process kiln, it was found that mudring formation
was caused by a large percentage of the coarse fraction in the kiln feed
which contained predominantly free silica. In another case, mudrings were
initiated by the kiln operator who did not excercise sufficiently tight con-
trol over the feed-end temperature, i.e, he allowed the temperature to vary
within a large range. Dust insufflation of the introduction of dust through
scoop feeders below the chain section can also be a primary cause for
mudring formations in wet-process kilns. Last but not least, mudrings
were also allowed to be formed simply by having the wrong type of chain-
system design or chain-link sixes.

References

1. Waddell, Joseph J. 1%2. Practical Quality Control for Concrete.


New YorkzMcGraw-Hill Book Co.
2. Bogue, RH. 1929. Industrial Engineering &m&y, Anal. Ed.
1:192.
3. Lea, F.M. and Parker, T.W. 1935. Building Research, Technical
Paper 16, London.
4. Kuehl,H.1929. Cement, 18, 833.
5. Lea, FM. and Desch, C.H. 1935. The Chemistry of Cement and
Concrete; Longmans, Green & Co.
12.

The Air Circuit in a Rotary Kiln

For a kiln operator, three variables that are related to air and gas flows
anywhere within the kiln system are of special interest. These are:

a) Volume and velocity


b) Temperature
c) Pressure

To determine the volume of a gas in a duct or flue, it is necessary to


measure both the temperature and the velocity. Velocity is determined by
means of a Pitot tube and differential pressure gauge. Other instruments
used for this purpose are: anemometers (when the velocities are less than 3
m/s), orifice disk or Venturi meters (when gases move at high velocity or
pressure through small pipes), or S-tubes that perform the same function
as Pitot tubes except that they are used for measurements of dust-entrained
gas streams. Strict rules apply and must be followed in order to obtain
meaningful and accurate results when velocity measurements are made in a
duct.
Pressure is measured and expressed by three different methods: static
pressure is measured at a right angle to the gas stream; impact pressure
(often referred to as total pressure) is measured by pointing the tube
directly into the center of the gas flow; and finally, the velocity pressure
(often referred to as dynamic pressure) is measured by subtracting the static
pressure from the impact pressure, as shown in Eq. (12-1).

Py=PI--Pps (12-1)
IMPACT STATIC
VELOCITY
PRESSURE PRESSURE
PRESSURE

Fig. 12.1 Pressure measurements.


THE AIR CIRCUR IN A ROTARY KILN

Instruments used for measuring pressures are: dial gauges, U-tubes, and
inclined draft gauges-the latter being the most sensitive and accurate of
the three.
After accurate pressure and temperature measurements have been taken
one can calculate the air or gas velocity by the following formula:

where:
vs = gas velocity per second
g = acceleration of gravity
PV = velocity pressure
Either English or metric units can be used in the above formula with the
appropriate use of the English or metric system gravitational constant.
The volume of gas passing through the duct is calculated by the
formula:

Q = 60AVs

where:
Q = volume of gas flow per minute
vs = average gas velocity (per second)
A = effective cross-sectional ama of the duct
Air movement through the kiln can be considered, for all practical
purposes, to take place uniformly from introduction of the air in the cooler
to final discharge at the stack. In accordance with the gas laws, however,
the physical state of the air undergoes changes in temperature, volume, and
pressure while traversing the kiln. The reasons for these changes is that
the chemical composition of the air itself changes in the process, and the
spaces through which the air travels are not uniform throughout the sys-
tem Figs. 12.2-12.4 show schematic diagrams of complete air circuits of
the various kiln systems in use in the cement industry.
For convenience, it is customary to divide this air flow into three suc-
cessive circuits in the following or&r:

a) The circuit for cooling of the clinker and introduction of


combustion air into the kiln. This is the cooler circuit,
b) The circuit in which combustion, calcination, and drying takes
place. This is the kiln circuit.
THE ROTARY CEMENT KILN

1 2

DISCHARGE KILN COOLING

Fig. 122 Air circuit on wet- and dly-process kilns-l and 2 am air inlets
into the undersrate chamber of the clinker cooler; 3 is primary air fan; 4, excess
coolerair stack; 5, kiln; 6 and 7, dust colkctors; 8. induced draft fan; 9, stack

c) The circuit in which the air and gases are released from the kiln
and pass through the heat recovery and dust-collecting units. This
is the discharge circuit.

In considering the air circuit in the kiln, it is necessary to look at all


three divisions as a whole because one leads naturally to the next. The
whole concept of air flow is one of a complete unit, as any adjustment or
change in one circuit will effect a change in the entire air circuit. Another
important factor is that it is almost impossible to avoid the entrance of so-
called cold parasite air into the kiln system This air enters through badly
closed openings around the kiln hood and around the back-end ducts. Not
only does this cut down the efficiency of the induced draft fan, but the cold
air entering through the hood and nose area reduces the fuel efficiency.
Drastic changes in the air flow take place when doors anywhere in the
system are left open for a prolonged length of time. Because this could
lead to upset kiln conditions, it is important to advise the operator in
advance when such an action is to be undertaken.

12.1 THE COOLER AIR CIRCUIT

The clinker cooler performs two functions: it must cool the hot clinker
discharged from the kiln and supply the kiln with the necessary air for
COOLlNG AIR IN
Fig. 12.4 Air circuit in precalciner kiln.
‘I-HE AIR CIRCUIT IN A ROTARY KILN

combustion. In doing so, valuable heat from the clinker is recuperated and
enters the kiln as hot secondary air. The most common clinker coolers in
use are:

a) inclined reciprocating grate coolers


b) planetary coolers
c) continuous traveling grate coolers
d) rotary coolers

Gn kihrs with large clinker outputs or special cooling problems, addi-


tional cooling of the clinker is often required after it discharges from these
coolers. This is being done by adding a so-called after-cooler to the above
units, the best known type of this kind being the G-cooler manufactured
by Claudius-Peters. Such after-coolers are also viable alternatives (instead
of installing completely new coolers) in cases when a kiln is being con-
verted from wet to dry, dry to preheater, or from preheater to precalciner
status, all of which produce a need for increased cooler capacity due to an-
ticipated higher clinker output rates. Unlike conventional coolers, there is
no direct contact between the clinker and the air in a G-cooler, hence, this
type of equipment is almost free of dust emissions. G-coolers provide the
final clinker cooling, receiving clinker at a temperature of around 350 C
(660 F) and discharging clinker at a temperature of approximately 100 C
(212 F).
The inclined reciprocating grate cookr is the most widely used type of
cooler today. This type of cooler derives its name from the reciprocating
(back-and-forth) movement of the rows of cooler grates that push the clin-
ker toward the discharge end. As many operators can attest to, this type of
cooler can not properly and efficiently perform its function when the
clinker is very fine as in the case where the kiln is in an upset condition.
Fig. 12.5 shows a schematic cross section of a closed-loop traveling grate
cooler. This type of reciprocating grate cooler has found some ap-
plications in wet-process kilns where combustion air requirements in the
kiln are usually very high. Circulating warm air from the lower section of
the cooler and using it for cooling air in the upper compartment lowers the
total excess air that has to be vented to the atmosphere and is thus bene-
ficial toward controlling the emissions from the cooler. However, this
system has its application limitations and is usually not suitable for large
dry and preheater kilns that are energy efficient and produce a large amount
of clinker per hour. Nevertheless, it is known that there is at least one
precalciner kiln that uses this kind of a system successfully in the United
THE ROTARY CEMENT KILN

Fig. 12.5 Cross section of closed loop traveling grate cooler, showing: 1,
kiln; 2, secondary air, 3, circulatory air; 4, hot air fan; 5, cold air fan; 6,
primary air fan; 7, excess cooler-air stack 8, dust collectors; 9, burner pipe; 10,
primary air; 11, clinker crusher; 12. traveling grates; A, B, and C are ambient
air inlets.

States. There is a limit to which air can be recycled within the cooler
compartments in order to maintain rapid quenching in the first cooler com-
partment. Coolers equipped with such a system often experience rapid
cooler fan blade wear when the recycled air contains a large amount of dust.
Cooler operating problems have become more pronounced as kilns
started to get larger and their output rates increased. Most grate coolers
require the use of a much larger amount of air for cooling the clinker than
what is needed for combustion in the kiln. Hence, modern coolers operate
with larger cooler excess-air volumes that are not usable for combustion in
the kiln. To prevent these heat losses, this excess air is usually used for
drying the kiln feed and/or coal and, in precalciner kilns, is transferred to
the flash furnace to be used as combustion air in the auxiliary firing unit.
Planetary coolers are cooling cylinders that are attached to the
circumference of the rotary kiln. There are no moving parts on these types
of coolers, no fans to control, and all the air that passes through these
tubes enters the kiln and is used for combustion. It’s the simplest system
for cooling clinker and there are no control functions except the rotational
THE AIR CIRCUIT IN A ROTARY KILN

speed of the kiln itself. Planetary coolers are equipped with lifters,
agitators, and blades to promote the heat exchange between the hot clinker
and the cold air. With all the air entering the kiln, there are no dust-
emission problems from these types of coolers. On larger units, noise
pollution can, however, become a problem from the clinker rattling in the
tubes. Planetary coolers require a larger I.D. fan than kilns of equal size
equipped with grate coolers but the total power requirements are much
lower when one considers all the air fans usually found on grate coolers.
The disadvantage of planetary coolers is that most will discharge the
clinker at a higher temperature than compatible grate coolers.
The continuous traveling grate coolers are similar to the reciprocating
grate coolers except that they use a continuous moving grate belt. Advan-
tages of this type of cooler are that individual worn or damaged plates can
be quickly replaced without needing to stop the kiln. Also, since the belt
travels from the hot to the colder part of the cooler and returns underneath
where it is being cooled by the cold air, overheating is less likely to occur
here than in reciprocating grates.
Rotary coolers were first used over 80 years ago, however, in order to
be efficient in today’s modem kilns they would have to be of such size as
to make them financially unjustifiable. There is some doubt that this type
of a cooler would ever make a comeback in the cement industry despite its
simplicity of operation.
Every cooling system must have an air balance to ensure proper cooling
of the clinker as well as proper combustion in the kiln. Any excess air
from the cooler not needed for combustion has to be vented to the atmos-
phere or must be used for other thermal work.
Efficient operation of a cooler requires that the minimum amount of air
with the lowest possible temperature be vented to the atmosphere while at
the same time the clinker is cooled to the temperature that ensures trouble-
free transport of the clinker to the storage area after it leaves the cooler.
Dampers and regulators must be so adjusted that no overheating of the
cooler equipment can take place, and the primary air and secondary air are
blended in the proper proportions when they enter the kiln. Temperature
of the secondary air entering the kiln should be as high as possible in order
to recoup a maximum amount of heat from the clinker, provided the cooler
and kiln nose are not permitted to become overheated.
On grate coolers, cooling is accomplished by forcing ambient air up-
ward through the grates and the bed of material. To obtain optimum
cooling it is important that the proper bed resistance is maintained. This
insures proper retention time for the clinker and efficient heat transfer. A s
THE ROTARY CEMENT KILN

the cooling air passes through the clinker bed and acquires a higher tem-
perature, its volume is expanded. As the air expands, more force is re-
quired to push this air through the bed. Since the material is hottest at the
inlet to the cooler, it follows that the cooling air expands the most here
and requires the highest pressure at this inlet side. Further down, the heat-
transfer rate becomes less, causing less expansion of the air and thus re-
quiring less pressure to get through the bed. This is the reason for
installing several compartments in these coolers and equipping each with a
fan of diminishing static-pressure rating. When this important factor is
overlooked in the initial design phase, or when the cooler is not properly
and efficiently compartmentalized, cooling efficiency of the unit as a whole
is sacrificed.
After the cooling air passes through the bed of clinker, the air needed for
combustion is drawn into the kiln as the so-called secondary air by the I.D.
fan or forced into the kiln as primary air by the primary air fan. The
remaining air is either vented to the atmosphere or used for drying coal or
raw feed. In the precalciner kiln, this excess air is transferred to the flash
furnace by means of the tertiary air duct.
It is a fundamental requirement that the air forced into the cooler is
applied to these areas where it can do the most useful thermal work. I n
thermodynamics, the best heat-transfer potentials take place where the
temperature differential is the greatest. It follows that it is necessary to
apply the maximum amount of air in the upper (hot) compartments and a
minimum amount in the lower (cold) compartments. This can be clearly
demonstrated by using the common heat-transfer Eq. (12-4):
Air Clinker
(1) (v.hO 02 - h) = (1) (sp.ht.) (T, - T2) W-4)

Inserting the appropriate values for the variables, this demonstrates that
one pound of ambient air has the capacity to lower the clinker temperature
in the fmt compartment from 1260 C (2300 F) down to 748 C (1378 F),
i.e., producing a temperature drop of 512 C (922 F) in the clinker. How-
ever, the same pound of ambient air, applied to the last cooler compart-
ment will be able to drop the temperature of a “colder” clinker of, e.g., 300
C (572 F) by only 13 C (23 F). In other words, the heat transfer in the
first compartment is 30 to 40 times more efficient than the heat transfer in
the lower compartment. This serves as a clear reminder to operators that
they should always first make sure that the air is properly applied in that
‘IlIE AIR CIRCUIT IN A ROTARY KILN

part of the cooler where it can do the most thermal work. This becomes
especially important in situations where a cooler exhibits very high excess-
air volumes that must be vented to the atmosphere. The same reasoning
must be applied when the clinker cooler consistently discharges
insufficiently cooled clinker that produces high finish-grinding tempera-
tures and could lead to quality-control problems in the cement. Many
cooler-operating problems have been, and can continue to be, solved by
applying this important rule of cooler operation. Damage to cooler com-
ponents often occurs when an operator neglects to recognize when the air
flow to individual cooler compartments is not properly adjusted to
prevailing operating conditions.
It is a recommended procedure that the plant engineering staff checks the
cooler air flows and establishes a cooler air balance at least once per year to
keep up-to-date information regarding the cooler’s efficiency.

12.2 THE ROTARY-KILN AIR CIRCUIT

The amount of air drawn through the rotary kiln is governed and
controlled soleZy by the induced draft fan (ID. fan) located at the feed end
of the kiln. The origin of the air going to the kiln is:

1) the hot secondary air emanating directly from the cooler,


2) the moderately heated primary air entering through the burner
pipe, and
3) the parasite air that enters through leaks and openings in the kiln-
hood area.

The ultimate objectives are: a) to induce a maximum amount of hot sec-


ondary air, b) to use only the minimum, safe amount of primary air that is
needed for fuel conveyance and flame control, and fmally c) to remove or
reduce the presence of parasite air. For every pound of cold air (e.g., 30 C)
drawn through these unnecessary in-leakage points in the hood, there is a
corresponding pound of valuable hot secondary air of around 800 C that
will not be put to useful work in the kiln. The end-result is that a higher
volume of excess cooler air has to be vented into the atmosphere. The
temperature of this air is also being increased resulting in the waste of
THE ROTARY CEMENT KILN

more valuable heat. The kiln operator must always try to stabilize the air
flow through the kiln. Unless the kiln &aft is controlled within
reasonable limits, a uniform firing rate and stable burning conditions can
never be obtained.
A certain definite amount of air is required for complete combustion in
the kiln, which means that close control over the fuel-to-air ratio has to be
exercised. Incomplete combustion, caused by the admission of too little
air, results in the loss of some unburned carbon because a portion of the
carbon in the fuel is burned to carbon monoxide (CO) instead of carbon
dioxide (CC+). To make sum that this does not take place a small amount
of excess air (approximately 5%) is introduced into the kiln. On the other
hand, too large an amount of excess air could lead to serious heat losses
because this air, not needed for combustion, raises the temperature of the
gases leaving the kiln at the rear. This rise in temperature in turn affects
the heat profile in the kiln, because of the change it causes in the drying,
calcination, and burning-zone conditions. Once a kiln operation is sta-
bilized, the following three basic rules must be observed in order to
maintain the kiln in such a condition:
a) Complete combustion conditions must prevail at all times while
a kiln is in operation, thus fuel and air rate changes can be
undertaken only to the extent that no formation of carbon
monoxide will result after the rates have been changed
b) The maximum amount of excess air present in the kiln should
not exceed the amount that will cause heat losses to the rear to
surpass the predetermined limit.
C) Temperatures of the air entering the kiln and leaving the kiln
should be held as nearly constant as possible in or&r not to upset
the heat profile in the kiln.

By weight, each unit of fuel burned (lb or kg) requires approximately


10.5 units (lb or kg) of combustion air. Air, at standard conditions, has a
specific volume of approximately 0.7735 m?kg (12.39 f&lb). Hence, to
burn 1 kg of fuel requires approximately 8.12 m3 of air for combustion in
the kiln. Likewise, burning 1 lb of fuel requires approximately 130 ft3 of
air at standard conditions. The gases produced during the combustion
process travel toward the rear of the kiln and are enriched with:

a) carbon dioxide evolved from the feed during calcination,


THE AIR CIRCUIT IN A ROTARY KILN

b) water from the evaporation of the moisture in the feed,


c) volatile constituents that have evolved (e.g., sulfur and alkalies),
and
d) dust evolved within the system

Clearly it stands to reason that a wet-process kiln, where all the calcination
(CO9 and all the evaporation of the moisture in the fed takes place within
the rotary kiln proper, has the highest gas volumes per unit weight of
clinker produced. Conversely, the precalciner kiln, where most of the cal-
cination is done in the preheater tower and no moisture is present in the
feed, has the lowest volume of kiln gases leaving the rotating kiln.
Gas-flow rates can be either calculated based on theoretical parameters or
actually measured by the aforementioned measuring methods. However,
actual measurements, particularly kiln exit-gas flows, will always be
suspect for accuracy unless strict adherence to established measuring rules
is practiced. It is also very difficult to establish the actual cross-sectional
area of the duct, at the point of measurement, because there usually are no
accurate means to determine the size of the coating buildup in the duct. It
is therefore good practice for an engineer to calculate the theoretical flow
rates as well as to ascertain that the actual measured values correspond
closely to the theoretical results.
In the following tables (Tables 12.1-12.4), kiln-air and gas-flow rates
of the four main kiln systems discussed in this book am shown. Oper-
ating condition (A) in Tables 12.1 and 12.2 refers to what is considered
efficient operations. Operating condition (B) in Tables 12.3 and 12.4 has
base data altered to show how the gas weights and volumes can change on
the same kilns as a result of less efficient operating conditions. These
tables have been developed on theoretical considerations only for the pur-
pose of demonstrating how changes in fuel and calcining rates can change
the gas-flow rates in the kiln.
Each individual type of kiln design has an established, by design,
pressure drop across the system Fans are thus designed to overcome this
pressure drop or resistance under normal operating conditons. Unfor-
tunately, these pressure drops do not remain constant since added restric-
tions in the form of buildups and rings can significantly increase this
pressure drop. Pressure-sensing devices that monitor pressure changes
within the system are a useful tool for operators to detect formation of
such restrictions. This is discussed in detail in the chapter on kiln control.
TABLE 12.1

WEIGHTS OF KILN AIR AND GAS FLOWS (typical efficient operation)


Wsinkts of Kiln Air and Gas-Flow Rates Opcm~& Cmdidon (A)

Base Data we+focess kiln w-ws~ Reheaterkiln Recalciner kiln

Fil&ll Metric English Metric English Metric English Metric


Fuel rate (IbIsht or kg/t) 3% 198 332 166 238.4 119.2 244 122
specific fuel consumption (MBtdt or kvkg) 4.95 5727 4.15 4821 2.98 3466 3.05 3547 3
Calcination dcne in kiln (%) 1 1 1 1 0.6 0.6 0.2 0.2
Fue1firedinflashthace(%) 0 0 0 0 0 0 0.5 0.5 T:

(values are in terms of lb&t clinker or k@t clinker) 2


4
Total air required for combustion in kiln 4158 2079 3486 1743 2503 1252 1281 641
secondary air 3326 1663 2789 1394 2253 1126 1153 516 8
l%imaryair 665 333 558 279 200 100 102
Infiltrated air at kiln hood 166 83 139 70 50 25 26 13
51 E
Combustion product in kiln 4633 2317 3884 1942 2789 1395 1427 714
Gases from the feed 2732 1366 11% 598 643 322 214 107
Total gasesleaving rotary kiln 7365 3683 5080 E
2540 3433 1716 1642 821
Infiltrated air at kiln bpck end 1395 698 1176 588 624 312 344 172
Total air reqired for combustion, flash furnace 0 0 0 0 0 0 1281 641
Combustion product in preheater tower 0 0 0 0 0 0 1427 714
Gases from the feed (preheater tower) 0 0 0 0 519 260 948 474

Total gases
leaving the kiln system 8760 4380 6256 3128 4576 2288 5642 2821
(i.e., going to I.D. fan)
TABLE 12.2
VOLUMES OF KILN AIR AND GAS FLOWS (typical efficient operation)
Volumes r$ Kiln Air and Gm-Flow Rates Operating Condition (A)
Base Data Wet-process kiln Dry-process kiln F-raderkiln precplciner kiln

English Metric English Metric English Metric English Metric

Fuel rate (wsht or kg/t)


specificfuelconsumpticln@mu/t a w/kg) 3 %
4.95 198
51.57 34.15
32 4186267 238.4
298 119.2
3466 244 122 k
3.05 3547 w
Calcination done in kiln (%) 1 1 1 0.6 00. 6 00.5
.2 0.2
Fuelfuedinfx4shhvnace(%) 0 0 01 0 0 0.5 G
T:
(in terms of SCF/sht aad Nm’/kg cl., standard comiitioa)
5
Total air required for combustion in kiln 51,518 1.61 43,192 1.35
1.08 31,015 0.97
0.87 15,872 0.50
0.45 5
secondary air 41214 1.29
34353 27,913 1434
6,911 2,481 MO
Infilkated air at kiln hood 2,061 0.06 1,728 0.05 620 0.02 317 0.01
Primaryair 8,243 0.26 0.22 0.08 0.04 i
Combustion product in kiln 59212 1.85 49,643 1.55 35,647 1.11 18,242 0.57
Gases from leaving
Total tbe feed kilo 32,210 1.01 12,881 0.40 6,775 0.21 2g1s 0.07
Infiltrated
gases
air at kiln
rotary
back end 91,423 285 62,524 1.95 45422 1.32 20,458 0.64
17284 0.54 14571 0.45 7,731 0.24 4262 0.13
Total air mquired for combustion, flash furnxe 0 0.00 0 0.00 0 0.00 15,603 0.49
Combustion product in flash furnace 0 0.00 0 0.00 0 0.00 1832 0.57
Gases from the feed (preheater tower) 0 0.00 0 0.00 5,336 0.17 9,793 0.31

Total
(i.e., gases
going leaving the
t o I.D. fao) kilo system 108,707 3.39 77,094 241 55,489 1.73 68,357 2.13
TABLE 12.3
WEIGHTS OF KILN AIR AND GAS FLOWS (assumed inefficient operation)
Weights qf Kiln Air ad Cm-Flow Rata Opmcing Candiria (B)

Base Data Wet-process kiln Dryprocess kihl Fkheater kihl Recalciner kihl

English Metric English Metric @l.ish Metric English Metric


cl
Fuel rate (lba.t or kg/t) 424 212 355.2 177.6 255.2 127.6 260.8 130.4
specitic fuel czoIlsulnF4ion (h4Bhdt o r w/kg) 5.3 6164 4.44 5164 3.19 3710 3.26 3192
calcination dale io kiln (%) I 1 1 1 0.85 0.85 0.35 0.35
Fuelfiredinflasbflmlace(%) 0 0 0 0 0 0 0.4 0.4

(v&m are io terms of lb/da clinker or k@t clinker)

Total air quid for combustion ia k i l n 4452 2226 3730 1865 2680 1340 1643 822
secon*pir 3562 1781 2984 1492 2412 1206 1479 739
Primarynir 712 3% 597 298 214 107 131 66
Infiltrated air at kiln h o o d 178 89 149 75 54 27 33 16

combustion product in kiln 4961 2480 4156 2078 2986 1493 1831 915
Gases from the feed 2732 1366 11% 598 911 4% 375 188
Total gases
leaving rotary kiln 7693 3846 5352 2676 3897 1949 2206 1103
hfiltratedairatkilnbskend 1550 775 1470 735 780 390 430 215
Total air required for combudm, flash furawe 0 0 0 0 0 0 1095 548
Combustion product in preheater towa 0 0 0 0 0 0 1221 610
Gases fmm the feed (pebeater tower) 0 0 0 0 251 125 787 394

Total gases
leaving the kiln system 9243 4621 6822 3411 4928 2464 5739 2870
(Le., going to I.D. fan)
TABLE 12.4
VOLUMES OF KILN AIR AND GAS FLOWS (assumed inefficient operation)
Vdwncs of Kiln Air and GM-Flow Rates operaing condilion (B)

Base Data w.qrocess kiln Dry-process kilo Reheater kiln Preoalciner kiln

English Metric English Metric English Metric English Metric


Foe1 rate (lb/At or kg/t) 424 212 355.2 171.6 255.2 127.6 260.8 130.4 22
P
specifk foe1 cotwmption (rvnwt or kJikg) 5.3 6164 4.44 5164 3.19 3710 3.26 3792 c!
Foelf~ioflashfomace.(%)
Calcination done in kiln (%) 01 01 01 0.7 0.35
0.4 0.5
01 0
0.85 0 0.2 E

(in terms of SCF/sb.t and NmQg CL, std. cooditioo) 5


Total air m&red for combustion in kiln 3390 z
55,160 1.72 46210 1.44 1.04 20,357 0.64
secondary air 44,128 1.38 36,968 1.15 29,880 0.93 18,322 0.57 P

MltlWdpiratkilUbood
Primalyair 2336
8,826 0.28
0.07 7,394
1,848 0.23
0.06 5656
664 0.08
0.02 1,629
407 0.05
0.01 5

combIlstion pduct in kiln 63,399 1.98 53,112 1.66 38,159 1.19 23,398 0.73 5
Gases hm tbe feed 32,210 1.01 15881 0.40 9,598 0.30 3,877 0.12
Total gases
leaving rotary kiln 95,609 298 65,993 206 47,757 1.49 27,275 0.85
Infiltr&edairatkilubackend 19205 0.60 18,213 0.57 9,664 0.30 5,328 0.17
Total air reqoired for combustion, flash fmnace 0. 0.00 0 0.00 0 0.00 13,341 0.42
Combustion product in flash lianace 0 0.00 0 0.00 2,5800 0.00 15399 0.49
Gases from tbe feed @reheater tower) 0 0.00 0 0.00 0.08 8,131 0.25

gases leaving the kiln system 114,814 3.58 awJ6 263 60,001 1.87 69,674 217
(i.e., going to I.D. fan)
THE ROTARY CEMENT KILN

12.3 THE DISCHARGE AIR CIRCUIT

Once the gases leave the kiln system (preheater tower), there remains
the task of removing the dust in these gases before they can be vented to
the atmosphere. Multiclones, cyclones, electrostatic precipitators, and/or
dust-bag houses are installed for this purpose. Without entering into a
detailed description of various dust-collecting systems, it can generally be
said that all these units cause an appreciable loss of pressure in the gas
stream, which makes a large capacity of the induced &aft fan a necessity.
The more dust-collecting units and the shorter the chimney or stack, the
more energy and fan capacity required to pull the gases through the kiln
system.
The operator should keep in mind that the I.D. fan is designed to handle
the hot kiln gases given off during normal operating conditions. If the fan
is called upon to move an equal volume of relatively cold gas, it could be
overloaded because the cold gas has a higher density (is heavier) than the
hot gas. For this reason, tables for maximum fan speeds at various exit-
gas temperatures must be furnished to operating personnel.
Although the discharge air cirucit is not directly related to the actual
production of cement, it nevertheless requires close attention from the
operator. An operator must focus his attention upon the temperature of
the gases at this location of the air circuit because either too high or low a
temperature can be harmful to the dust-collecting equipment. The mois-
ture and the sulfur content of the gases require that this temperature be
controlled within a predetermined range. If the temperature falls below the
dewpoint of the gases, the moisture can precipitate out in the baghouse or
electrostatic precipitator, causing plug-up problems and chemical attack
due to corrosion. Filter bags also have a given maximum-service tem-
perature above which these bags can be damaged. Most baghouses require
that the gases passing through do not exceed 285 C (545 F). The dew-
point of the gases varies depending on local prevailing conditions.
Experience has shown that when the dust-collector inlet temperature is held
above 160 C (320 F) there exists little danger of moisture precipitating out
in the collector. In summary, it can be stated that most kiln systems have
to operate the dust-collector inlet temperature within the range of 160-285
C (320-545 F) to prevent operating problems in the dust collector. It is
important to remember that these temperatures apply to the dust-collector
inlet and not to the kiln exit or preheater-tower exit-gas temperature. The
THE AIR CIRCUlT IN A ROTARY KILN 173

reader might question the aforementioned dewpoint temperature for it is


true that the dewpoint for kiln gases is normally between 50-90 C
(120-195 F). Why then the adherence to the higher temperature given
above? The answer is infiltration in the dust collector itself. Kiln gases
are being further cooled inside these units by as much as 60 C (140 F)
which can bring these gases close to the dewpoint even when they enter at
a much higher temperature.
Another factor that has to be considered in respect to the discharge air
circuit is the air inleakage between the kiln exit and the dust collector.
This inleakage should be held to a minimum by effective sealing of all
unnecessary openings. These openings diminish the induced draft fan (I.D.
fan) capacity, require extra power to drive this fan, and could drop the gas
temperature below the previously discussed, harmful dewpoint. It can be
calculated that 25% infiltrated air at the discharge circuit can increase the
horsepower requirements of the ID. fan by an approximately equal amount
of 2% which is a significant inefficiency in the system Sealing doors
and openings within the gas ducts is normally not difficult but, obtaining
an effective kiln back-end seal is usually associated with high installation
and maintenance costs. Nevertheless, these expenditures for effective seals
can be financially justified by the resulting reduced horsepower require-
ments of the fan.
The amount of air infiltration at the discharge circuit can be measured
by extracting and analyzing simultaneously the gases at the kiln exit and
the dust-collector inlet. The percent infiltrated air is then calculated by Eq.
(12-5):

I = oxf-0x0 xl00
20.9 - OXf
where:
I = percent of infiltrated air (by volume)
O X , = percent O2 at dust-collector inlet
OX, = percent 0, at kiln discharge.
Any other two sampling points can be selected in the system, such as dust-
collector inlet/dust-collector outlet, to determine infiltration.
13.

Movement of Material Through the Kiln

13.1 KILN LOADING

The feed, as it advances through the kiln from the rear, through the
several zones, and on to the cooler, does not move at a uniform speed, nor
does it move in a straight line along the axis of the kiln. In the dehy-
dration reach of the kiln behind the calcining zone, the feed first enters the
chain section in which chains, attached loosely to the interior of the kiln,
serve to agitate the feed to facilitate heat transfer. The speed with which
the feed advances through this section is largely influenced by the chain
pattern and density. The agitation and tumbling action cause considerable
mixing of the feed constituents. Upon leaving the chain section the feed is
raised part way up the arch as the kiln turns, following a path perpen-
dicular to the kiln axis. At a certain point gravity causes the mass to slide
down the kiln shell in a forward direction, and the movement is then
repeated. In this manner the feed advances in a zig-zag path, each rise and
slide advancing all parts of the feed a few inches. In contrast to the chain
section, very little mixing takes place in the feed bed in this area.
The feed now enters the calcining zone, and the zig-zag course dimin-
ishes progressively as the feed bed travels down the kiln, because, with the
evolution of carbon dioxide gas, the feed bed becomes partly fluidized and
travels more rapidly, flowing in a manner somewhat akin to the flow of
water, as well as continuing the rise-and-slide advancement. A slowdown
occurs next in the burning zone. The feed, now transformed into a semi-
liquid state, becomes sticky, starts to form clinkers, and undergoes a cas-
cading action similar to that in the dehydration zone.
Naturally, there are irregularities that occur in this movement of feed
through the kiln. Sometimes, when viewing the burning zone through the
MOVEMENT OF MATERIAL THROUGH THE KILN

observation port, one can observe an onrush of dust waves that flush with
considerable speed into the burning zone. These waves, moving much
faster than the regularly advancing material bed, originate in the calcining
zone because of erratic calcination resulting from some irregularity in kiln
operation which caused the bed to become fluidized. They are much more
frequent in smalldiameter kilns with steep slopes than in large-diameter
kilns with smaller inclination.
The opposite condition sometimes occurs in which material is retained
for a prolonged period of time in a certain area of the kiln, caused by ring
formation in the back of the burning zone or less frequently in the chain
section. When any such ring breaks loose the kiln operator is confronted
with a so-called “push” in which all the feed retained by the ring enters the
burning zone along with fragments of the broken ring.
In both cases these reactions cause an uneven bed depth and constituue
difficult operating conditions which can cause the kiln to go into an upset.
Such upsets can occur in different magnitudes. Usually only small
adjustments to kiln operation are necessary to maintain uniformity of
operation of the kiln. At other times the condition may be so severe as to
require more drastic measures, even to the extent of taking emergency
procedures such as shutting down the kiln to protect the equipment. For
these reasons it is useless to attempt to explain the steps necessary to
counteract changes in feed-bed movement. Only experience andknowledge
of each individual kiln will enable a kiln operator to know what kind of
counteracting procedure should be carried out for any given condition. It
should be emphasized, however, that it is extremely important to consider
the after-effects an upset can have in the feed bed in the areas behind the
burning zone. If the operator is content to take care only of the material in
the burning zone during either a push or a light load, without anticipating
what changes have taken place in the rear of the kiln, the kiln might be
permitted to overreact to the opposite extreme. This could lead the kiln
into a cycling condition that would require hours or even days to eliminate.
The mentioned zig-zag movement of the feed in the calcining zone is
not conducive to good heat exchange between the gases and the feed. For
lack of a tumbling action, the center of the feed bed experiences little
contact with the gases whereas the particles on the surface get maximum
exposure and thus maximum heat exchange. This action causes an uneven
calcination within the cross section of the feed bed. Only shortly before
the feed enters the burning zone will the increased tumbling action cause
the “cold” core of the bed to be exposed to the gases. At that time, feed
THEi ROTARY CEMENT KILN

will then undergo its final and rapid calcination. This action is especially
noticeable and pronounced in long wet- and dry-process kilns. Uneven
calcination of the feed bed can often give an operator the false impression
that the kiln-feed composition is too difficult to burn, i.e., the laboratory
is mistakenly blamed for having prepared a tough-burning mix.
On wet- and dry-process kilns, lifters and cam linings are designed to
promote mixing of the feed bed in the caking zone and thus obtain a
more uniform and better heat exchange for the bed as a whole. Many
plants, however, experience higher and sometimes unacceptable dust gen-
eration as a result of these heat exchanger installations. A new type of
lifter, called a disperser, has been tried on a few dry and wet kilns with
good initial results. The intention of this disperser is to accomplish a
stirring of the feed bed without causing the feed to be lifted into the gas
stream. The dispersers are designed so that the feed slides across the uphill
side and tumbles over a ledge on the downhill side. Operating data on
these kilns has shown improved kiln stability, lower specific fuel con-
sumption, and in one instance an increase in kiln output. Most sig-
nificantly, the internal dust cycle and dust-waste rates do not increase
noticeably after these dispersers are installed On the negative side it must
be mentioned that these dispersers are difficult to maintain and show a high
wear and spall rate.
Under normal conditions, it is possible to compute, with reasonable
accuracy, the time required for the feed to move through the kiln. Gibbs1
proposes this equation:
11.4L
T=-
NDS

in which

T = traveltime(min)
L = length of kihr (ft) or (m)
N = speed of kiln (revolutions per hour)
D = internal diameter of kiln (ft) or (m)
S = slope of kiln (ft/ft) or (m/m)
The constant 11.4 applies to cement kilns and is obtained from the
relationship:

constant = 1.77 * 4 $
MOVEMENT OF MATERIAL THROUGH THE KILN

where:
$ is the average angle of repose for the material bed (assumed to be
between 35 and 43 degrees for cement kilns). However, visual observation
of the burning zone shows this angle to be more in the magnitude of 50
degrees. In the calcining zone, this angle is also considerably less due to
the partial fluidization of the bed. Tests performed by the author on a
rotating cylinder (1.0 r-pm), having a smooth lining surface, showed an
angle of repose of less than 20 degrees.
In Table 13.1 the retention time, feed-bed velocity, and percent kiln
loading have been calculated to demonstrate, in theory, how these variables
can change as the feed bed travels through the different zones down the
kiln. The angles of repose and kiln-feed densities have been estimated to
present, as closely as possible, actual kiln conditions.
Percent loading of the kiln is calculated as follows:
a) English system:
loq1.3888Wj)
d
% loading =

- A
0 t

b) Metric system:

lOO(O.69444wJ)

% loading = d

where:
w = kiln output (sh.tons/day) or (metric tons/day)
f = feed factor (lb feed/lb clinker)
d = feed density (lb/fts) or (kg/ms)
1 = kiln length (ft) or (m)
t = retention time (min)
A = effective cross-sectional area inside kiln (ft2) or (m2)
The percent loading of the kiln is a factor that is of considerable interest
TABLE 13.1
DRY-PROCESS KILN RETENTION TIME AND KILN’LOADING
Kiln Data:

Kiln length: 525 Kiln speed: 72


Kiln I.D.: 14.50 Retention time (Gibbs formula): 1 5 7
Slope: 0.0365

Total and E
Preheat 10% 20% 40% 80% Burning average ;;I
ZOW cakined cakined calcined calcined ZO?lf? VdUS 2
Feed factor 1.58 1.522 8
1.464 1.348 1.116 1.00 1.383
2928 26% 2232 2000 2765
Feed 3 80 40 4 9
Length(lb)/clinker(t)
of zone (ft) 3210600 60 50 40 95 525
Estimated angle of repose 45 27 24 21 18 50 36.42
Inside diameter (ft) 14.5 14.5 14.5 14 13.25 12.8 14.05
Estimated density (lb/f@) 75 73 70 65 60 92 75.10

Retention time (min) 65 20 14 12 9 37 157


Feed velocity (fthin) 3.06 3.95 4.19 4.32 4.42 2.56 3.3
Loading (a) 10 7.7 7.3 7.5 7.3 7.9 8.6
MOVEMENT OF MATERIAL THROUGH THE KILN

to the engineer and operator. It is also a subject that can generate


considerable disagreement among operators for one might prefer to run the
kiln at higher speed and lower percent loading whereas another might prefer
the opposite, namely, slow speed with a corresponding deeper bed. Oper-
ators can vary this loading by adjustment to:

a) the kiln speed, or


b) the feed rate.

Most kilns operate at 610% loading while kilns with a steep slope tend
to operate at the higher end and kilns with shallower slopes on the lower
end of this scale. The optimum setting can usually only be found by trial
and error. To find this optimum setting, a kiln is usually operated for a
few days at various percent loadings and its effect on kiln operating stabil-
ity and efficiency investigated. The apparent “ideal” setting will then
become the standard for all future operations on that kiln. More discus-
sions on this subject follow later when kiln-speed and feed-rate controls are
individually discussed.

13.2 KILN CAPACITY

A multitude of formulas have been advanced by various writers per-


taining to the determination of the kiln capacity. Results obtained from
these calculations can vary to a considerable extent. The choice of
formulas narrows down to the question of what is meant by the word
capacity. A kiln manufacturer, for example, will state the kiln capacity in
terms of what is considered attainable, reasonable output rates, i.e.,
conservative. It is not unusual to exceed this stated capacity by as much
as 10-U% for this author has consistently worked kilns that outproduced
their rated capacity. Often, too, the operators are heard mentioning that a
given kiln is capable of producing a lot more clinker but that either the
I.D. fan, the cooler, or the raw-grinding capacity is a bottleneck. Then
there are numerous cases where operators might be of the opinion,
sometimes rightfully so, that the kiln is being “force” fired, i.e., being
crammed with feed to such an extent that it causes undue operating
problems.
THE ROTARY CEMENT KILN

G. Martin1 ma& a strong point in 1937 when he determined that the


kiln capacity is not a function of the overall kiln volume but that the kiln
diameter is the governing factor in determining how much a kiln is capable
of producing. At the time when he wrote his book, there were no pre-
calciner kilns in existence to give him the proof that his reasoning was
correct. The formulas he advanced for determining the kiln capacity have
also become outdated with the installation of induced draft fans on kilns
because his calculations were based on natural kiln drafts.
From an operator’s viewpoint, kiln capacity is that limit above which
gas velocities become so high that kiln feed is easily lifted into the gas
stream, i.e., the internal dust cycle becomes excessively large and starts to
interfere with the stability of the operation. These conditions are reached
when the kiln-gas velocity exceeds a speed of approximately 9 m/s (30
ft/s). A kiln that operates in the calcining zone with a higher velocity
than indicated above is considered to be “force fired.” Since this is the so-
called threshold limit, kilns usually are operated at output rates that are
5-10% below the velocity-calculated capacity.
The kiln-gas velocity is calculated by the well-known formula:
Velocity = volume/area
and is thus influenced by the following:
a) the effective cross-sectional area of the kiln
l kiln diameter inside lining
l percent loading of kiln

l inserts such as chains, lifters

l coating and ring formation


b) the volume of gas passing through this cross section
0 fuel firing rate (exclude fuel to precalciner)
l gases emanating from the feed

l temperature of the gas

l amount of excess air in kiln

With the help of a computer the capacity tables (Tables 13.2-13.8)


have been developed to show the threshold (limit) output rates for various
kiln systems. The tables are printed in both the English and Metric
systems of units, hence, care has to be taken to read the appropriate tables.
To simplify computations, the following constants have been used in the
development of these tables:
1 Ci. Martin, Chemical engineering and hmnodynamics applied to rolary cement kilns.
London: Technical Press Ltd.
MOVEMENT OF MATERIAL THROUGH THE KILN

English Metric

Fuel heating value 12,600 BhVlb 29,288 kJ/kg


Combustion gas 11.6 lb/lb coal 11.6 kg/kg coal
(@ 5% excess air)
Gas from feed 0.534 lb/lb clinker 0.534 kg/kg clinker
(@ 100% calcination)
Gas density (std. cond.) 0.09 lb/ft3 1.44 kgINm3
Loading in calciniig zone (%) 9% 9%
Gas temperature (dry and wet kiln) 1600 F 871 C
@reheater kiln) 1900 F 1038 c
(precalciner kiln) 2150 F 1177 c

These tables can be used to answer the questions if a given kiln is being
“force fired” or if the output is considerably below its capacity. If a kiln
output is higher than indicated in the table, and if this kiln is plagued with
instability, an effort should be made to lower the production rate. On the
other hand, when a given kiln consistently produces less than 80% of the
indicated capacity, efforts should be undertaken to raise the output provided
the auxiliary equipment can handle this increase.
TABLE 13.2
WET-PROCESS KILN CAPACITY
English units (sh.t./h)

9% loading in calcining zone

Kiln I.D. MBtu/t clinker

4.8 5.0 5.2 5.4 5.6 5.8 6.0 6.2 8z


2
11 10 33 60 .. 62 29.3 28.3 23 73 .. 52 232.3
6.7 23 51 .. 93 23 50 .. 25 229.6
4.5 2
35.4 34.3
12 43.5 42.1 40.8 39.6 38.4 37.3 36.2 35.3 i2

11 43 55 91 .. 31 57.3
49.4 45 75 .. 95 45 63 .. 49 52.3
45.1 54 03 .. 88 44 92 .. 35 44 81 .. 04 E
15 68.0 65.8 63.8 61.8 60.0 58.3 56.6 55.1
16 77.4 74.9 72.5 70.3 68.3 66.3 64.4 62.7
17 87.4 84.6 81.9 79.4 77.1 74.8 72.7 70.8
18 98.0 94.8 91.8 89.0 86.4 83.9 81.6 79.3
19 109.2 105.6 102.3 99.2 96.2 93.5 90.9 88.4
20 121.0 117.0 113.4 109.9 106.6 103.6 100.7 98.0
21 133.4 129.0 125.0 121.2 117.6 114.2 111.0 108.0
Metric units (metric t/h clinker)
5
9% loading in calcining zone
d
Kiln I.D. kJ/kg clinker
8
5579 5811 6044 6276 6508 6741 6973 7206 g
3.049 27.4 26.5 25.7 24.9 24.2 23.5 22.8 22.2 F
3.354 33.2 32.1 31.1 30.2 29.3 28.4 27.6 26.9 ;;f
3.659 39.5 38.2 37.0 35.9 34.8 33.8 32.9 32.0
3.963 46.4 44.9 43.4 42.1 40.9 39.7 38.6 37.5 E
4.268 53.8 52.0 50.4 48.9 47.4 46.0 44.8 43.5
2
4573 61.7 59.7 57.8 56.1 54.4 52.9 51.4 50.0
4.878 70.2 68.0 65.8 63.8 61.9 60.1 58.5 56.9 g
5.183 79.3 76.7 74.3 72.0 69.9 67.9 66.0 64.2
g
5.488 88.9 86.0 83.3 80.8 78.4 76.1 74.0 72.0
ti
5.793 99.0 95.8 92.8 90.0 87.3 84.8 82.4 80.2 Ia
6.098 109.7 106.2 102.8 99.7 %.8 94.0 91.3 88.9
6.402 121.0 117.1 113.4 109.9 106.7 103.6 100.7 98.0 $
I
TABLE 13.3
DRY-PROCESS KILN CAPACITY
English units (sh.tih)
9% loading in calcining zone

Kiln I.D. MBNt clinker

3.8 3.9 4 4.1 4.2 4.3 4.4 4.5

10 36.3 35.6 34.9 34.3 33.6 33.0 32.4 31.8


11 44.0 43.1 42.3 41.5 40.7 39.9 39.2 38.5
12 52.3 51.3 50.3 49.3 48.4 47.5 46.7 45.9
13 61.4 60.2 59.0 57.9 56.8 55.8 54.8 53.8
14 71.2 69.8 68.5 67.2 65.9 64.7 63.5 62.4
15 81.8 80.1 78.6 77.1 75.7 74.3 72.9 71.6
16 93.0 91.2 89.4 87.7 86.1 84.5 83.0 81.5
17 105.0 102.9 100.9 99.0 97.2 95.4 93.7 92.0
18 117.7 115.4 113.2 111.0 108.9 107.0 105.0 103.2
19 131.2 128.6 126.1 123.7 121.4 119.2 117.0 115.0
20 145.3 142.5 139.7 137.1 134.5 132.0 129.7 127.4
21 160.2 157.1 154.0 151.1 148.3 145.6 143.0 140.4
Metric units (metric t/h)
8
9% loading in calcining zone
2
Kiln I.D. kJ/kg clinker
3
4416 4533 4649 4765 4881 4998 5114 5230
9
3.049 33.0 32.3 31.7 31.1 30.5 29.9 29.4 28.9
F
3.354 39.9 39.1 38.3 37.6 36.9 36.2 35.6 35.0
3.659 47.5 46.5 45.6 44.8 43.9 43.1 42.3 41.6 l!l
3.963 55.7 54.6 53.6 52.5 51.6 50.6 49.7 48.8
F
4.268 64.6 63.3 62.1 60.9 59.8 58.7 57.6 56.6
4.573 74.2 72.7 71.3 69.9 68.6 67.4 66.2 65.0 2
4.878 84.4 82.7 81.1 79.6 78.1 76.7 75.3 74.0 8
5.183 95.3 93.4 91.6 89.8 88.2 86.5 85.0 83.5
5.488 106.8 104.7 102.7 100.7 98.8 97.0 95.3 93.6 g
5.793 119.0 116.7 114.4 112.2 110.1 108.1 106.2 104.3
6.098 131.9 129.3 126.8 124.3 122.0 119.8 117.6 115.6 E
6.402 145.4 142.5 139.7 137.1 134.5 132.1 129.7 127.4
TABLE 13.4
SUSPENSION PREHEATER KILN CAPACITY (20% calcined in cyclones)
English units (sh.t/h)

80% calcination done in kiln


9% loading in calcining zone

Kiln I.D. MBNt clinker


zz
2.8 2.9 3 3.1 3.2 3.3 3.4 3.5 2
s
10 42.2 41.1 40.0 39.1 38.1 37.2 36.3 35.5
11 51.1 49.7 48.5 47.3 46.1 45.0 44.0 43.0 i2
12 60.8 59.2 57.7 56.2 54.9 53.6 52.3 51.2 5
13 71.3 69.5 67.7 66.0 64.4 62.9 61.4 60.0 5
14 82.7 80.5 78.5 76.5 74.7 72.9 71.2 69.6
15 94.9 92.5 90.1 87.9 85.7 83.7 81.8 79.9
16 108.0 105.2 102.5 100.0 97.6 95.3 93.1 91.0
17 122.0 118.8 115.7 112.9 110.1 107.5 105.0 102.7
18 136.7 133.2 129.8 126.5 123.5 120.6 117.8 115.1
19 152.3 148.4 144.6 141.0 137.6 134.3 131.2 128.3
20 168.8 164.4 160.2 156.2 152.4 148.8 145.4 142.1
21 186.1 181.2 176.6 172.2 168.1 164.1 160.3 156.7
Metric units (ndric t/h)
8
80% calcination done in kiln
9% loading in calcining mne 2
Kiln I.D. kJ/kg clinker 8

3835 3952 4068 8


3254 3370 3487 3603 3719

3.049 38.3 37.3 36.3 35.4 34.6 33.8 33.0 32.2


s
3.354 46.3 45.1 44.0 42.9 41.8 40.8 39.9 39.0
3.659 55.1 53.7 52.3 51.0 49.8 48.6 47.5 46.4
E
3.%3 64.7 63.0 61.4 59.9 58.4 57.0 55.7 545
4.268 75.0 73.1 71.2 69.4 67.8 66.2 64.6 63.2 El
4.573 86.1 83.9 81.7 79.7 77.8 75.9 74.2 72.5 g
4.878 98.0 95.4 93.0 90.7 88.5 86.4 84.4 82.5
5.183 110.6 107.7 105.0 102.4 99.9 97.6 95.3 93.1 g
5.488 124.0 120.8 117.7 114.8 112.0 109.4 106.8 104.4
5.793 138.2 134.6 131.2 124.8 121.9 119.0 116.4 8
127.9
6.098 153.1 149.1 145.3 141.7 138.3 135.0 131.9 128.9 E
6.402 168.8 164.4 160.2 156.3 152.5 148.9 145.4 142.1 2
TABLE 13.5
SUSPENSION PREHEATER KILN CAPACITY (40% calcined in cyclones)
English unifs (sh.t./h)
60% calcbtion done in kiln
9% loading in calcining zone

Kiln I.D. MBtuk clinker


g
2.8 2.9 3.0 3.1 3.2 3.3 3.4 3.5 ;;r
10 45.0 43.7 42.6 41.4 40.4 39.4 38.4 37.5 2
11 54.4 52.9 51.5 50.1 48.9 47.6 46.5 45.4 9

11 32 76.0 76 33 .. 90 761.3
1.9 75 09 .. 07 658.1
8.2 65 66 .. 57 65 45 .. 93 63.4
54.0 B
64.8
14 88.2 85.7 83.4 812 79.1 77.2 75.3 73.5
15 101.2 98.4 95.8 93.2 90.9 88.6 86.4 84.4 E
16 115.2 112.0 109.0 106.1 103.4 100.8 98.3 96.0 2
17 130.0 126.4 123.0 119.8 116.7 113.8 111.0 108.4
18 145.8 141.7 137.9 134.3 130.8 127.6 124.4 121.5
19 162.4 157.9 153.7 149.6 145.8 142.1 138.7 135.3
20 180.0 175.0 170.3 165.8 161.5 157.5 153.6 150.0
21 198.4 192.9 187.7 182.8 178.1 173.6 169.4 165.3
Metric units (metric th)
60% calcination done in kiln
,9% loading in calcining zone

Kiln I.D. Id/kg clinker


3254 3370 3487 3719 3835 3952 4068 si
bd
3.049 40.8 39.7 38.6 37.6 36.6 35.7 34.8 34.0
3.354 49.4 48.0 46.7 45.5 44.3 43.2 42.2 41.2
3.659 58.8 57.2 55.6 54.1 52.7 51.4 50.2 49.0
3.%3 69.0 67.1 65.3 63.5 61.9 60.4 58.9 57.5 F
4.268 80.0 77.8 75.7 73.7 71.8 70.0 68.3
66.7 2
4.573 91.9 89.3 86.9 84.6 82.4 80.4 78.4 76.5
89.2 87.1 Fs
4.878 104.5 101.6 98.9 96.2 93.8 91.4
5.183 118.0 114.7 111.6 108.7 105.9 103.2 100.7 98.3
110.2 g
5.488 132.3 128.6 125.1 121.8 118.7 115.7 112.9
5.793 147.4 143.3 139.4 135.7 132.2 128.9 125.8 122.8 8
6.098 163.3 158.8 154.5 150.4 146.5 142.9 139.4 136.1 Lx!
6.402 180.0 175.0 170.3 165.8 161.5 157.5 153.7 150.0 F
z
TABLE 13.6
PRECALCINER KILN CAPACITY (50% fuel to precalciner)
English units (sh.t/h)

0.1% cakination done in kiln


0.5% of total heat input into kiln
9% loading in calcining zone

K&I I.D. MBtu/t input into kiln t5


2
1.35 1.40 1.45 1.50 1.55 1.60 1.65 1.70
z
10 107.3 103.8 100.5 97.4 94.4 91.7 89.1 86.6 2
11 129.8 125.6 121.6 117.8 114.3 110.9 107.8 104.8
12 154.5 149.4 144.7 140.2 136.0 132.0 128.3 124.7
150.5 146.4 4
13 181.4 175.4 169.8 164.5 159.6 154.9
14 210.3 203.4 1%.9 190.8 185.1 179.7 174.6 169.8
15 241.5 233.5 226.0 219.0 212.5 206.3 200.4 194.9
16 274.7 265.7 257.2 249.2 241.7 234.7 228.0 221.8
17 310.1 299.9 290.3 281.3 272.9 264.9 257.4 250.4
18 347.7 336.2 325.5 315.4 305.9 297.0 288.6 280.7
19 387.4 374.6 362.7 351.4 340.9 331.0 321.6 312.7
20 429.2 415.1 401.8 389.4 377.7 366.7 356.3 346.5
21 473.2 457.6 443.0 429.3 416.4 404.3 392.9 382.0
Metric units (metric t/h)

0.1% calcination done in kiln


0.5% of total heat input into kiln
9% loading in calcining zone

Kiln I.D. kJ/kg input into kiln

1569 1627 1685 1743 1801 1860 1918 1976

3.049 97.4 94.1 91.1 88.3 85.7 83.2 80.8 78.6


3.354 117.8 113.9 110.3 106.9 103.7 100.6 97.8 95.1
3.659 140.2 135.6 131.2 127.2 123.4 119.8 116.4 113.2
3.963 164.5 159.1 154.0 149.3 144.8 140.6 136.6 132.8
4.268 190.8 1845 178.6 173.1 167.9 163.0 158.4 154.0
4.573 219.0 211.8 205.1 198.7 192.7 187.1 181.8 176.8
4.878 249.2 241.0 233.3 226.1 219.3 212.9 206.9 201.2
5.183 281.3 272.1 263.4 255.2 247.6 240.4 233.6 227.1
5.488 315.4 305.0 295.3 286.1 277.6 269.5 261.8 254.6
5.793 351.4 339.9 329.0 318.8 309.3 300.2 291.7 283.7
6.098 389.4 376.6 364.5 353.3 342.7 332.7 323.3 314.4
6.402 429.3 415.2 401.9 389.5 377.8 366.8 356.4 346.6
TABLE 13.7
PRECALCINER KILN CAPACITY (45% fuel to precalciner)
English unifs (sh.Vh)

0.2% calcination done in kiln


0.55% of total heat input into kilu
9% loading in calcining zone

Kiln I.D. MBtu/t input into kiln


E
1.485 1.540 1 s95 1.650 1.705 1.760 1.815 1.870 2
10 91.6 88.8 86.1 83.6 81.2 79.0 76.9 74.8 z

11 12 110.9
131.9 107.4
127.8 104.2
124.0 101.1
120.4 98.3
117.0 95.6
113.7 93.0
110.7 90.6
107.8 B
13 154.8 150.0 145.5 141.3 137.3 133.5 129.9 126.5 2
14 179.6 174.0 168.8 163.8 159.2 154.8 150.6 146.7
15 206.2 199.8 193.7 188.1 182.7 177.7 172.9 168.4
16 234.6 227.3 220.4 214.0 207.9 202.2 196.7 191.6
17 264.8 256.6 248.8 241.6 234.7 228.2 222.1 216.3
18 296.9 287.6 279.0 270.8 263.1 255.9 249.0 242.5
19 330.8 320.5 310.8 301.8 293.2 285.1 277.4 270.2
20 366.5 355.1 344.4 334.4 324.9 315.9 307.4 299.4
21 404.1 391.5 379.7 368.6 358.2 348.3 338.9 330.1
Metric units (metric t/h)

0.2% calcination done in kiln


0.55% of total heat input into kiln 5
9% loading in calcining zone s

Kilo I.D. kJ/kg input into kiln 9


1726 1790 1854 1918 1982 2046 2109 2173 52

3.049 83.1 80.5 78.1 75.8 73.7 71.6 69.7 67.9 5


3.354 100.6 97.5 94.5 91.8 89.2 86.7 84.4 82.2 3
3.659 119.7 116.0 112.5 109.2 106.1 103.2 100.4 97.8
3.963 140.5 136.1 132.0 128.2 124.5 121.1 117.8 114.7 E
4.268 162.9 157.9 153.1 148.6 144.4 140.4 136.7 133.1 5!
4.573 187.0 181.2 175.8 170.6 165.8 161.2 156.9 152.8 8
4.878 212.8 206.2 200.0 194.1 188.6 183.4 178.5 173.8
5.183 240.2 232.8 225.8 219.2 212.9 207.1 201.5 196.2 z
5.488 269.3 261.0 253.1 245.7 238.7 232.1 225.9 220.0
5.793 300.1 290.8 282.0 273.8 266.0 258.6 251.7 245.1 8
6.098 332.5 322.2 312.5 303.3 294.7 286.6 278.9 271.6 E
6.402 366.6 355.2 344.5 334.4 324.9 316.0 307.5 299.4 2
TABLE 13.8
PRECALCINER KILN CAPACITY (40% fuel to precalciner)

0.3% calcination done in kihl


0.6% of total heat input into kiln
9% loading in calcining zone

Kih~ I.D. MBtuh input into kiln

1.62 1.68 1.74 1.80 1.86 1.92 1.98 2.04

10 79.9 77.6 75.3 73.2 71.2 69.4 67.6 65.9


11 %.7 93.9 91.2 88.6 86.2 83.9 81.8 79.7
12 115.1 111.7 108.5 105.5 102.6 99.9 97.3 94.9
13 135.1 131.1 127.3 123.8 120.4 117.2 114.2 111.3
14 156.7 152.0 147.7 143.5 139.6 136.0 132.4 129.1
15 179.9 174.5 1695 164.8 160.3 156.1 152.0 148.2
16 204.6 198.6 192.9 1875 182.4 177.6 173.0 168.6
17 231.0 224.2 217.7 211.7 205.9 200.5 195.3 190.4
18 259.0 251.3 244.1 237.3 230.8 224.7 218.9 213.4
19 288.6 280.0 272.0 264.4 257.2 250.4 243.9 237.8
20 319.7 310.3 301.4 293.0 285.0 277.5 270.3 263.5
21 352.5 342.1 332.3 323.0 314.2 305.9 298.0 290.5
luerric imirs (metric t/h)
0.3% calcination done in kiln
0.6% of total heat input into kiln
9% loading in calcining z&me

Kiln I.D. kJ/kg input into kiln

1883 1953 2022 2092 2162 2231 2301 2371 +l


68.4 66.4 64.6 62.9 61.3 59.8 ii4
F
3.049 72.5 70.4
3.354 87.7 85.2 82.7 80.4 78.2 76.1 74.2 72.3
3.659 104.4 101.3 98.4 95.7 93.1 90.6 88.3 86.1
115.5 112.3 109.2 106.3 103.6 101.0 F
3.963 122.6 118.9
4.268 142.1 137.9 134.0 130.2 126.7 123.3 120.2 117.1 s!
4.573 163.2 158.3 153.8 1495 145.4 141.6 137.9 134.5 i5
4.878 185.6 180.2
5.183 209.6 203.4 175.0
197.5 170.1
192.0 165.5
186.8 161.1
181.9 156.9
177.2 153.0
172.7 3
5.488 235.0 228.0 2215 215.3 209.4 203.9 198.6 193.6
5.793 246.7 239.9 233.3 227.2 221.3 215.8 ii
261.8 254.0
6.098 290.1 281.5 273.4 265.8 258.5 251.7 245.2 239.1 E
6.402 319.8 310.3 301.4 293.0 285.0 277.5 270.4 263.6 5
TABLE 13.9
PRECALCINER KILN CAPACITY (30% fuel to precalciner)
EngIish units (sh.ti)

0.4% calcination done in kiln


0.7% of total heat input into kiln
9% loading in calcining zone

Kiln I.D. MENU/t input into kiln z5


1.89 l.% 2.03 2.10 2.17 2.24 2.31 2.38 ;;t
2
10 66.8 64.9 63.1 61.4 59.7 58.2 56.7 55.3
3
11 80.9 78.5 76.3 74.2 72.3 70.4 68.6 66.9
12 96.2 93.5 90.8 88.4 86.0 83.8 81.7 79.7
13 112.9 109.7 106.6 103.7 100.9 98.3 95.8 93.5 B
14 131.0 127.2 123.6 120.3 117.1 114.0 111.2 108.4
15 150.4 146.0 141.9 138.1 134.4 130.9 127.6 124.5
16 171.1 166.1 161.5 157.1 152.9 148.9 145.2 141.6
17 193.1 187.6 182.3 177.3 172.6 168.1 163.9 159.9
18 216.5 210.3 204.4 198.8 193.5 188.5 183.7 179.2
19 241.3 234.3 227.7 221.5 215.6 210.0 204.7 199.7
20 267.3 259.6 252.3 245.4 238.9 232.7 226.8 221.3
21 294.7 286.2 278.2 270.6 263.4 256.6 250.1 243.9
Metric units (metric t/h)

0.4% calcination done in kiln


0.7% of total heat input into kiln
9% loading in calcining zane

Kiln I.D. kmg input into kiln

2197 2278 2359 2441 2522 2603 2685 2766 6i

3.049 60.6 58.9 57.2 55.7 54.2 52.8 51.4 50.2 5


3.354 73.4 71.2 69.2 67.4 65.6 63.9 62.3 60.7 i4
3.659 87.3 84.8 82.4 80.2 78.0 76.0 74.1 72.3 F
3.963 102.5 99.5 %.7 94.1 91.6 89.2 86.9 84.8
4.268 118.8 115.4 112.2 109.1 106.2 103.4 100.8 98.4 ii
4.573 136.4 132.5 128.8 125.2 121.9 118.8 115.8 112.9 ti
4.878 155.2 150.7 146.5 142.5 138.7 135.1
5.183 175.2 170.2 165.4 160.9 156.6 152.5 131.7
148.7 128.5
145.0 z
5.488 196.4 190.8 185.4 180.3 175.6 171.0
5.793 218.9 212.5 206.6 200.9 195.6 190.5 166.7
185.7 162.6
181.2 2
6.098 242.5 235.5 228.9 222.6 216.7 211.1 205.8 200.7 E
6.402 267.4 259.6 252.4 245.5 238.9 232.8 226.9 221.3 E
Part II

Kiln Operating Procedures


14.

Kiln Operating and Control Methods

Methods of kiln control vary from plant to plant, and even between
different operators, because each operator is apt to have his own ideas as to
how to proceed when confronted with any given situation. Kiln control,
however, must be a continuous around-the-clock matter, hence it is
necessary for all operators on all shifts to operate the kiln in the same
manner. This in turn means that all operators should be trained in the
same principles of kiln operation.
There is nothing more destructive to operating stability than the
changing of controller setpoints en masse during shift changes. Unless an
emergency exists or an obvious change must be ma&, it is not possible
for an operator to assess the need for controller-setting changes in the first
few minutes after coming on duty. A difference in settings from the
previous day is not necessarily an indication to change these settings back
to where they were 24 h before. An operator should allow 20 min at the
onset of the shift to fust observe the process before deciding that a change
is indeed warranted.

14.1 CLINKER-BURNING TECHNIQUES

There are three common techniques for burning clinker in a rotary kiln:

a) Maintain a constant kiln speed, and vary the fuel rate to


counteract temperature changes in the burning zone.
THE ROTARY CEMENT KILN

b) Maintain a constant fuel rate, and vary the kiln speed to hold the
burning-zone temperature at the desired level.
C) Vary the kiln speed, the fuel rate, or both, to maintain the desired
burning-zone temperature.

These techniques have one error in common; they show concern only
for the burning-zone temperature. Unfortunately, many kiln operators
think that this is good enough, reasoning that, as long as good clinker is
produced, what more is necessary? The fallacy of this reasoning lies in the
fact that ideal stable kiln conditions can be obtained faster and more
economically when equal consideration is given to all zones in the kiln and
not the burning zone alone. Drying and calcining of the feed must be
considered before one can consider making clinker. The process of clinker
burning, and therefore the process of rotary-kiln control, starts not at the
place where the feed enters the burning zone, but at the point where the
feed enters the kiln. In preheater and precalciner kilns this applies to the
point where the feed is given to the top stage of the preheater cyclones.
The technique described in this book can be summarized as follows:

VARY THE KILN SPEED, THE FUEL RATE, AND THE


INDUCED DRAFT-FAN SPEED IN ANY COMBINATION TO
MAINTAIN THE PROPER BURNING-ZONE TEMPERATURE
AND A CONSTANT KXN BACK-END TEMPERATURE FOR A
GIVEN FEED RATE.

This is sometimes referred to as “burning a kiln from the rear,” the


most reliable technique that fulfills the four fundamental rules of clinker
burning which, in order of priority, are:

a) Protection of the equipment and personnel at all times.


b) Production of well-burned clinker.
c) Continuous stable kiln operation.
d) Maximum production with maximum fuel efficiency.

a) Protection of Equipment and Personnel. Safe operation must have


first priority at all times; the operator should never allow himself to
bypass this rule. Red grates in the cooler, red spots on the kiln shell, or
overheated chains can do much more damage than the production of a few
barrels of bad clinker, if bringing the equipment trouble under control
KILN OPERATING AND CONTROL METHODS

would happen to result in the production of a small amount of poor


clinker. It is usually during times of unusual and severe upset of the
operating conditions that an operator will most likely forget this fun-
damental rule. This is because of the attention understandably being given
in such instances to leading the kiln back to normal operating conditions.
But, it is precisely during these “tight” moments that the shift in attention
toward the safety of the equipment and fellow workers must be made. The
majority of accidents happen when the kiln is in an unusual operating
condition such as during kiln starts, stops, and upsets. Safety around a
rotary kiln being an important and integral part of kiln operation is covered
in greater details in Chapter 26.
b) Production of a Well-Burned Clinker. A well-burned clinker is a
clinker that is neither underburned nor overburned, has been properly
cooled, and possesses the correct free-lime content and the desired liter
weight.
c) Continuous Stable Kiln Operation. Continuous operation should
always have priority over maximum production. More can be gained by
having emphasis on continuous operation rather than by pushing the kiln
to peak production at the expense of periodical kiln upsets. Stable kiln
operation is the key to long refractory life, high fuel efficiency, and
uniform quality clinker.
The term “stable kiln condition” means the condition in which only
very small changes, or no changes at all, have to be made to the control
variables to hold the kiln in a state of equilibrium It is the kiln operator’s
duty to obtain stable conditions and not be satisfied until this goal has
been achieved. The capability of a kiln operator is measured not so much
by the length of time a stable kiln condition is maintained but by how
well kiln upsets are handled, and how much skill is used in leading the
kiln back to stable conditions after an upset.
Stable operation is also identified by the recording charts, most note-
worthy being the kiln speed, and burning-zone and back-end temperatures
(all deviating very little over a long period of time). Upset conditions are
marked by large changes in kiln speed and, when these speed changes occur
in frequent intervals, e.g., every 2 h, the kiln is in a cycling condition.
Working a kiln out of a cycle, i.e., breaking a cycle, is the most chal-
lenging task a kiln operator faces (see Chapter 25). This is the time when
kiln burning becomes more of an art than a routine. Such kiln upsets can
happen not only on dry- and wet-process kilns but, although less severe
and less frequent, also on preheater and precalciner kilns. Even kilns with
THE ROTARY CEMENT KILN

sophisticated computer and automatic control are not immune to these


cycling conditions.
d) Maximum Production with Maximum Fuel Effciency. Only after
the three previous requirements have been met can one try to raise produc-
stion and start to concentrate on the details of fuel efficiency. Any move
made in this direction should be done very slowly in order to avoid
upsetting the prevailing stable kiln condition. Large kilns usually operate
better in the upper range of their rated production rate, hence production
increases can help to stabilize the operation if the kiln has been operating
somewhat below its rated capacity.

14.2 MANUAL CONTROL

Controlling a kiln in the conventional manner, as was done on most


kilns twenty years ago, consisted of observing the operation and making a
manual adjustment on the control panel based upon that observation. One
of the many possible rotary kiln control functions is looking into the
burning zone and noticing a change in temperature which requires an ad-
justment in the fuel rate. Experience tells the kiln operator that he can
neither turn the fuel valve wide open nor fully closed, but must base the
required fuel change on how much the temperature has changed. A small
variation in temperature requires only a small adjustment, while a large
temperature change necessitates a correspondingly large modification in the
amount of fuel admitted into the kiln.
The next factor to take into consideration is the limit to which the
burning-zone temperature can drop or increase without requiring more
drastic action than adjustment to the fuel rate itself. Too much fuel
addition, although justified by the existing burning-zone temperature,
could lead later on to dangerous overheating or incomplete combustion.
Likewise, cutting the fuel rate back futher than an established limit might
lead the kiln into a severe upset condition. Before any adjustment is made,
the operator must consider the limit to which the burning-zone temperature
and the fuel rate can be permitted to go without leading the kiln into an
upset.
After an adjustment has been ma& to the fuel rate, the next question is
how long the fuel rate can be left at the new setting until it has to be
adjusted again. A certain time will transpire from the moment the fuel rate
is adjusted until the burning zone reacts to this change and returns to its
KILN OPERATING AND CONTROL METHODS

target temperature again. However, when the temperature has reacted and
starts to turn in the other direction, at a given point another fuel-rate
adjustment must be made so that the temperature does not overshoot its
target. In other words, the temperature should not move from one extreme
to another but should return and level off at the target. This is the essence
of manual process control and the thought process required of the operator.
By now it should be obvious that the skill and experience of the
operator play an all-important part in this type of control. His judgment
and decisions am the main factors governing the degree of stability the kiln
operation will obtain. If one considers the fact that an operator often has
to exercise control over not one but sometimes two and more kilns at the
same time and that each kiln has a multitude of controls such as described
above, it is understandable that many of these fine aspects of control can
be overlooked. Every kiln operator, regardless of how long he has been on
the job, knows that an error in judgment can sometimes be made or a
wrong decision can be reached. Often a decision could not be made strictly
according “to the book” at the time it had to be made because of some
uncertainty in the process not apparent to the operator. The only thing
known at the time could have been that some adjustment was necessary.
This adjustment may be shown on the strip-chart recording a few hours
later to be completely contrary to what should have been done. Nobody
feels worse than the operator himself when such evidence shows up on the
strip chart Thus, it is easy to say afterwards that something different
should have been done, but it is far more difficult to make the right
decision at the time when some uncertainty existed and in an immediate
decision had to be made.
Manual control undoubtedly demands a heavy work load from the
operator because most of these manual adjustments are repetitious in
nature. Fortunately, the cement industry has experienced a phenomenal
revolution in automatic process control in the past 20 years. This has
made the manually controlled kilns more of an exception and it is expected
that, a few years hence, they will become a rarity.
A process as complicated as the operation of a rotary kiln necessitates a
multitude of instruments and controllers. It would be impossible to con-
trol a kiln without the help of instruments. Not too long ago, kiln con-
trol was limited to measuring temperatures, pressures, and flow rates,
transmitting the information to a recorder on the fling floor, then leaving
actual control of the variables in the hands of the kiln operator. It was not
unusual to observe an operator going from burner hood porthole to control
panel and back again, making almost continuous adjustments to the
THE ROTARY CEMENT KILN

controls. Understandably, operators became uneasy when the reactions


didn’t come out exactly as expected, as individual human error played a
large part in this method of operating.

14.3 AUTOMATIC CONTROL

Today’s kilns are not only equipped with more and better controllers,
but the computer is also slowly taking over the controls of entire kiln
systems. Automation has become the standard in the cement industry,
performing the work the kiln operator had to do himself in the past.
Concurrent with this revolution, there also appeared new terminologies,
better known as computer jargon that the kiln operator must become
familiar with. During the early stage, when automation was still in its
infancy, there were numerous workers that showed apprehension and
sometimes outright objection to these developments. Advances made in
control technology over the past few years have mostly eliminated this
initial apprehension. Computer systems have been “humanized,” made
more simple to understand, and control capabilities have become so
reliable and advanced that even the most experienced kiln operators accept
them as a blessing. It has definite advantages over pacing the burner floor
for eight hours especially during cold winter and hot summer days. Con-
trol-room operators also have an ally in the computer engineer who can be
of assistance when operating problems have to be resolved.
It would defeat the purpose of this book to discuss in details the
numerous types of computer systems that are being used today in the
cement industry. Little benefit could be derived by discussing the control
concepts of any particular kiln, because each kiln has, to some extent, its
own design and its own automatic control idiosyncracy.
INSTRUMEYTATION

15.

Instrumentation

15.1 AUTOMATIC CONTROLLERS

Having automatic controllers and computers to handle part or all of the


kiln operation requires that the operator acquires a basic knowledge of the
workings of an automatic controller. There is a specific need to familiarize
an operator with this subject, for nobody else has a better opportunity to
observe the results of an automatic controller in actual operation than he
does. Automatic process control has a better chance of succeeding and will
yield better results if the experience, observations, and suggestions of the
operators are considered when the automatic control is initiated or adjusted.
In this chapter, some of the apparently difficult-to-understand principles
of process control are explained in simple terms, using as an example the
control of the temperature in the burning zone. The fact that repeated
reference is made to the burning-zone temperature and the fuel rate does not
mean that the described fundamentals are valuable for this example alone.
These fundamentals are applicable almost universally to any kind of
process control and the proper relationship between the example given here
and a specific process under investigation should be easily recognized.
Regardless of whether a kiln is manual, automatic, or fully computer-
controlled, trends of the more important process variables are either re-
corded on a recording chart or displayed on the CRT. Fig. 15.2 is a
portion of a strip recorder chart similar to the one being examined in Fig.
15.3. This chart also shows the effect of “process noise,” the normal
momentary fluctuations (but not changes) in the variable being recorded.
Line A is virtually free of process noise, but Line B suffers from con-
spicuous noise.
A recorder, by itself, only records what the input variable tells it to
record. In other words, a recorder merely takes a visible record of a process
Fig. 15.1 Instrumentation for a modem uxncnt mill consists of scores of
recorders, contmIlcrs, and indicators, as shown in this view of a portion of a
central control room. (YE F0xbro C O.)
INSTRUMENTATION

Fig. 15.2 Two variables are traced on this strip chart. Line A represents the
grate speed in the cooler. Speed is quite uniform and the curve is free of process
noise. Line B, percentage of oxygen in the kiln exit gas, shows a large
increase in oxygen starting about 10115 am, reaching a maximum at 11:25, then
decreasing. Maximum span for oxygen content is from O-56; span reaches
from 15-70% of chart (0.75-3.52 oxygen). Oxygen curve shows considerable
process noise.

variable such as pressure, flow, temperature, or speed.


Under automatic control a controller in response to input signals,
makes changes in the output variable, or control means, a process function
that maintains the input variable within certin established limits. These
changes or adjustments to the output variable can also be recorded (Fig.
15.4). For example, in response to variations in burning-zone tem-
perature, the controller makes changes in the rate of fuel consumption.
An integral part of the controller system is the trmsmitter. Each of the
two flow transmitters shown in Fig. 15.5 sends electrical signals to the
recording controller which records and controls the flow in accordance with
instructions that were programmed into the controller. A schematic
diagram of instrumentation for a kiln and cooler is given in Fig. 15.6.
Automatic controllers are installed for the specific purpose of
improving overall control of a process and eliminating costly human errors
THE ROTARY CEMENT KILN

Fig. 15.3 A strip chart from the process recorder in the instrument panel is
being examined by the operator. Two controllers are located immediately to the
right of the recorder. (The Jhboro CO.)

that can occur when a process is manually controlled. When a controller


meets these objectives, the work of the operator is made considerably
easier (see Fig. 15.1).
The easy way out for an operator, when things do not go according to
plan, is to take control manually and say that the controller is no good.
The better approach, and the only one which can lead to satisfactory end
result, is to find out why the controller does not do the job as it originally
was designed to do. A serious effort in this direction will in most cases
yield an answer.
When manual control is to be replaced by an automatic control, reliable
instruments (in the following example, a pyrometer) have to be found and
properly installed in order that a true measurement of the process variable
can be obtained. Obtaining this correct value or process trend is an
absolute requirement in automatic control. If this is not done, the effort to
achieve automatic control is doomed to fail. Simply stated, the question
becomes: How is it possible to control any variable, for example, a level
INSTRUMENTATION

CLOSED OPEN

Fig. 15.4 Range on this outpt variable (fuel rate) chart, being 40% of
maximum range, reaches from the 20% to the 60% lines of the chart.

in a tank or bin, if the instrument that measures this level does not work
or measure properly? The answer to this question is obvious and needs no
further clarification.
In order that this discussion is not drawn into fields and details beyond
the control of the operators, assume that all other possible obstructions
and important factors for the installation of an automatic control system
have been taken care of and all necessary equipment for this type of control
has been installed.
THE ROTARY CEMENT KILN

Fig. 15.5 Each of these flow transmitters sends mcasurcment signals to a


recorder or controller. (7%e Foxboro C O.)

15.2 TUNING (PROGRAMMING)

The first and most important step is the tuning of the controller.
Tuning simply means making the controller do what we want it to do, by
setting a set point and tuning certain dials to get an optimum response
from the controller. When these settings are properly made, the controller
will make the correct response to signals received from the process
equipment. When these settings are incorrectly adjusted, regardless of how
well and how elaborately the system has been designed, the end results in
control will not be satisfactory. An automatic controller or a computer is
ELECTRONIC INSTRUMENTATION

CEMENT KILN AND CLINKER COOLER

Fig. 15.6 Typical instrumentation for kiln and cooler. (The I;oxboro C O.)
THE ROTARY CEMENT KILN

capable of doing only what it has been instructed to do-nothing more and
nothing less. A mistake on automatic control is far more serious than a
mistake on a process that is manually controlled. While an operator in
manual control might make a mistake once in a while, an automatic
controller will repeat the same mistake time after time without end, if
programmed improperly.
In or&r to arrive at the proper setting for controller action on any
process loop that is to be controlled, there am several variables that must
be known before the control can be established,
First is the spun, which is the range that can be tolerated between low
and high values of the input variable. It is expressed as a percentage of the
maximum span. Second is the maximum span, which is the largest pos-
sible variation of the input variable that can be measured on the instru-
ments. Maximum span is normally considered to be 100% of the chart.
These relationships are illustrated in Fig. 15.7. Third, the operating range
is the desired percent range between minimum and maximum allowable
limits in the value of the output variable (Fig. 15.4). Usually it is not
possible to operate the output device over its entire range (e.g., a valve
from fully open to fully closed) so an operating range must be selected that
will provide the most effkient control.
In any process-control system, the sefpoinr is the ideal value of the
input variable about which the process is controlled. Because many
recorder charts are graduated in percent of chart,* it is customary to locate
the setpoint at a point expressed by a certain percentage of the chart,
usually 5&70%. For example, consider a linear recorder so designed and
operated that its maximum span of 100% represents a temperature range
from 2000” at zero percentage of chart to 3200” at 100% of chart
Maximum span of this chart (O-100%) represents a range of 1200°.
Assuming that 2600° is the ideal temperature at which a good clinker can
be produced without overbuming or underbuming, then 50% would be
selected as the setpoint. Span would then be selected to cover whatever
extent is allowable for the temperature to vary. Fig. 15.8 shows the
relationship between percentage points on the chart and actual values being
charted, in this case temperature. The operator in Fig. 15.9 is manually
adjusting the setpoint on a recorder.
A convenient conversion table can be developed for the fuel rate (see
Table 15.1). By using this table, the operator can determine the quantity

* The tmnd now is towards charts graduated in actual engineering units lather than in
percent.
INSTRUMENTATION

Flg. 15.7 Relationship of spen and maximum span on a typical input


variable chart for burning-zone temperature. for which span, reaching from the
15% line to the 85% line, is 70% of maximum span.

of heat input at any fuel flow rate for either coal, gas, or oil at any per-
centage point on the recorder chart. This information is useful in deter-
mining whether the fuel is being used efficiently at any certain clinker
production rate. It also tells the operator what equivalent setting to use
when changing from one fuel to another.
THE ROTARY CEMENT KILN

- - -

- - - -

d -

0 20 4 0 6 0 00 100%
2000* 2mOe 3200.

P%%T

Fig. 15.8 Relationship between percentage on the chart and actual values, in
this case temperature, is shown.

TABLE 15.1
FUEL RATE CONVERSION*
Gas rale Oil rate
Percent BBI Percent BB1
Chalt fWh MBwh clinker/h chart gam MBtu/h clinker/h

3 69.6 74.9 96 20 480 72.5 93


12 138.8 149.5 192 42 1008 152.2 195
15 155.2 167.2 214 46 1104 166.7 214
20 179.2 193.0 247 53 1272 192.1 246
26 204.0 219.7 282 61 1464 221.1 283
36 240.0 258.5 331 71 1704 257.3 330
40 253.2 272.7 350 75 1800 271.8 348
100 400.0 430.8 552 100 2400 362.4 465

*This table was extracted from a larger table for certain specific fuels and equipment.
Similar tables can be computed for any pIapt.

In the discussion of manual control an example was cited in which the


factors to be considered were described by a kiln operator before and after an
adjustment in fuel rate was carried out. For automatic control to be
INSTRUMENTATION

Fig. 15.9 The operator is adjusting the setpoint dial. (The Foxboro Co.)

successful, the controller must be programmed in such a way that it con-


trols the temperature and adjusts the fuel rate in a similar fashion.
Most controllers have either one, two, or three action controls by which
this can be accomplished. They are: proportional, reset, and rule action.
In most applications proportional plus reset functions (two-mode) are
utilized. When rate is utilized as a compensation for time lag in
measurement, it is generally in a controller with proportional, reset, and
rate function (three-mode).
Proportional Action. In a proportional-action controller, the output
variable, or control means, is set in a specific relationship to the input
variable, or process variable. The mechanics of setting the controller is
simply an adjustment of a dial on the controller that shows the respective
settings. A large range of settings is available depending on the particular
instrument in use.
THE ROTARY CEMENT KILN

An important concept involved in proportional-action control is the


proportional band, this being the ratio of desired span (input) to operating
range (output) expressed as a percentage. That is
Span
proportional band* = 100 (15-l)
Operating range
If the output variable is allowed to move through an operating range of
100% then the proportional band equals the span. If, however, the output
operating range is less than 100, then the proportional band varies, as
shown in Table 15.2 by some values selected at random At a constant
span for the input variable, the proportional band is increased to decrease
the output range. Obviously, identical proportional bands are possible
under completely different conditions in process control.

TABLE 15.2
VARIATION IN PROPORTIONAL BAND
Input variable I Output variable
Example Burning-zone temperature Fuel late Percent
number percent chalt percent chart propoltional
band
Max. Min. Span Max. Min. Range

1 95 25 70 28 24 4 1750
2 90 30 60 30 22 8 750
3 80 35 45 30 24 6 750
4 85 35 50 32 20 12 420
5 83 38 45 34 19 15 300
6 80 40 40 35 18 17 235
7 75 45 30 38 16 22 136
8 90 30 60 56 12 44 136
9 70 50 20 40 14 26 77
10 67 52 15 41 13 28 54
11 80 20 60 100 0 100 60

Again referring to the example in which burning-zone temperature is


being controlled, Fig. 15.7 illustrates the concept of span and maximum
span for the input variable under control, in this case burning-zone
temperature. Span, arbitrarily set at 70% of maximum span, is selected as

+ Note that “gain” is the inverse of pqxnkmal band.


INSTRUMENTATION

being the range of temperature that can be controlled by fuel-rate adjust-


ments alone. In other words, the temperature will be permitted to vary
over this range. Whenever the temperature starts to vary from the set-
point, the control is actuated and the fuel valve opens or closes as required
depending on whether the temperature is below or above the setpoint.
This is illustrated in Fig. 15.4 which shows the fuel-valve action over an
operating range of 40% in response to the temperature input shown in Fig.
15.7.
Selection of proportional band settings can have a significant effect on
the reaction of the output device. Fig. 15.10 is a stylized drawing of a
recorder chart showing the burning-zone temperature deviating from set-
point. The output variable response is large or small depending on the
proportional band. The relationship can be computed by the equation:
lOOt?
m=- (15-2)
P
in which
m = percent chart response of output &vice
e = percent chart change in input variable
p = proportional band setting.

Example: What percentage will the fuel rate change on the chart when
the burning-zone temperature changes 20% on the chart, at a controller
proportional band setting of 300?
100x20
m = ___ = 6.7
300
Answer: The fuel rate will change 6.7% on the chart
Proportional bands are changed according to the response of the process.
For example, a water-level recorder might be set once and never changed.
Kiln burning-zone temperature control, on the other hand, is more
dynamic, and requires daily checking by an experienced person who is
qualified to make any adjustments that may become necessary. Proper
evaluation will avoid overcontrol, a common failing of the inexperienced
operator, who is apt to make two or more simultaneous changes on a
controller. Changes should be made one at a time, each change evaluated
before going to the next one (if another one is actually necessary).
The serious shortcoming of proportional action control is the failure of
the system to control at the setpoint in the process under control. This
THE! ROTARY CEMENT KILN

OUTPUT INPUT
FUEL RATE B.Z. TEMP.

0 20 40 60 80 100 %

Fig. 15.10 PROPORTIONAL, ACTION. Fuel-valve response to a change in


burning-zone te-mperahue at various proportional band settings.

means that proportional action does not necessarily bring the process
variable (input variable) back to the desired setpoint when it deviates. All
it does is to move the output variable in proportion to a change in the
input variable.
Reset Action. Most processes require that the process variable be held
at or returned to a specific setpoint for optimum and efficient operation.
We have learned that proportional action does not consider the setpoint and
will not necessarily return the process variable exactly to the setpoint once
the process has had an upset. Another aspect of process control unac-
counted for in proportional action is the process reaction time or lag.
Often several minutes can pass after an adjustment has been made to the
INSTFUJMENTATION

output variable before the process reacts to the change. With reset action,
the means are available to return the process back to the setpoint and to
account for possible delayed process reactions.
Reset action causes the output device to change at a rate proportional to
the deviation of the input variable from the setpoint. How often this
action is carried out (repeated) is governed by the reset-time setting on the
controller. Reset time can be expressed as minutes per repeat or seconds
per repeat For example with a reset-time setting of one minute the action
is repeated once every minute. With a setting of 22 min, the reset action
would be repeated once every 22 min. Reset action enters into the control
as long as the process variable is deviating from the setpoint. As the
departure of the input variable from the setpoint decreases, the amount the
output device corrects is decreased, during the course of any correction.
When the process variable levels off at the setpoint, reset action will stop
and does not enter the process again until the variable starts to move away
from the setpoint.
This holds true in our exampIe of burning-zone temperature control.
When reset action is applied, the percentage change in fuel feed (fuel-valve
movement) is identical to the percentage deviation of the burning-zone
temperature from the setpoint. Added weight must be given to the
setpoint in this instance because too high or too low a temperature can not
only damage the quality of the clinker, but can also damage the coating and
refractory as well as impair the overall operation of the kiln itself. With a
change in the fuel rate the burning-zone temperature does not return to an
acceptable level immediately thereafter. As pointed out previously, several
minutes can pass before the temperature starts to reverese its deviation and
move towards the setpoint. This delayed reaction can be observed in other
areas of a rotary-kiln system. For example, when the bed-grate drive speed
in the clinker cooler is increased in order to reduce the bed thickness, a
considerable length of time can pass until the thickness is at the desired
level.
The manner in which changes in reset time affect the process is shown
in Fig. 15.11, which demonstrates that the output variable or control
means (fuel response in this example) can be delayed or accelerated in
response to departure of the input variable (burning-zone temperature) from
the setpoint. Reset times (in minutes) am indicated by the figures adjacent
to the fuel-rate response lines.
Reset rate, sometimes used instead of reset tune, is merely the
reciprocal of reset time. For example, if reset time is 2, then reset rate is
0.5 repeats per second.
THE ROTARY CEMENT KILN

RESPONSE B.Z. TEMP.

I
0 20 40 50 60 80 100
SET
POINT

Fig. 15.11 RESET ACTION. A long rcsct time results in smaller output
response compared with a short reset time.

Reset action response is practically instantaneous whenever the


temperature moves away from the setpoint, i.e., it reacts immediately to
the temperature difference or error and starts the correction. This response
of reset action has been mathematically expressed.

Percentage movement 1 percentage deviation (15-3)


of output variable = reSet time x of input variable

Thus, when the temperature suddenly changes 15% and the reset time is 5
min, the fuel-valve response would then be:
INSTRUMENTATION

1
x 15 = 3.0 percentage points per minute
3
Converting into response per second yields 3.0/60 = 0.05 percentage
points per second. Another example: The temperature changes 15% but
this time the reset time is 20 rnin:
1
- x 15 = 0.75 percentage points per minute or
20 0.0125 percentage points per second

From this formula another important equation can be established:


percentage deviation of input variable from setpoint
Reset time =
percentage movement of output variable in a given time
Remember to use the same time units in the result as stated in the output
variable movement. That is, if the output variable movement is expressed
in seconds, then reset time is likewise expressed in seconds.
Reset action eliminates cycling conditions in the process caused by the
earlier-mentioned reaction time after an adjustment has been made to the
control device. In setting reset time on the controller, one usually starts
with a long reset time and progressively reduces the time setting until the
cycling condition stops.
Derivufive Action. The third and last control action available on some
controllers is derivative action, or rate action. Derivative action is usually
combined with proportional, or proportional plus reset action, and is
almost never used alone to control a process. It is used as a compensation
for time lag or inertia of the measured variable.
Derivative action causes the output device to respond proportionally to
the rate of change of the input variable when it is changing. In our
example, when the burning-zone temperature moves rapidly in either
direction, the fuel-rate response will be correspondingly fast, and when the
temperature moves slowly, the fuel-rate response will be correspondingly
slow. In essence, derivative action is the same as the procedure most kiln
operators commonly use when they control a rotary kiln manually. When
they look into the burning zone and notice a large, rapid temperature
change, they “hit” the kiln with a large change in fuel rate. When the
temperature changes only slowly, they “nurse” the kiln with small
adjustments in the fuel rate. Derivative action diminishes to zero when the
burning-zone temperature stabilizes. One must remember that derivative
action, like proportional action, does not consider the setpoint. In other
THE ROTARY CEMENT KILN

OFFSET

:;
2
F 1 \
3

FUEL RATE B.Z. TEMP.

Fig. 15.12 DERIVATIVE ACTION. In Curve 1, the derivative time (fuel-rate


response) is relatively slow, in response to a slow rate of change of the input
variable @uming-zone temperature); Curves 2 and 3 represent progressively
shorter derivative times.

words, derivative action ceases as soon as there is no more change in the


input variable, e.g., the burning-zone temperature, and does not consider
whether the temperature levels off at the setpoint or at some other
temperature.
Derivative action is adjusted on the controller as a function of time just
as in reset action. Thus, one refers to derivative time whenever an
adjustment has to be made on the controller. In Fig. 15.12, Curve 1 repre-
sents a long derivative time (slow response), and Curve 3 represents a
short derivative time (fast response).
INSTFWh4ENTATION 225

FUEL RATE B.Z.TEMf?

rn

Fig. 15.13 Combined proportional plus reset action is shown.

Propor&al Plus Reset Action. From the previous discussion, it can


be seen that proportional action sets the output device proportional to the
change in the input variable, and reset action changes the output device at a
rate proportional to the deviation of the input variable from the setpoint.
Fig. 15.13 shows how these two actions can be combined, again using
burning-zone control as an example. When no reset action exists (reset
time set at infinity), the fuel-rate response would be identical to the propor-
tional action. However, with the two control methods combined, the fuel
rate is influenced by the reset action. Thus, at time increment 7 the
temperature has leveled off, consequently proportional action has stopped
because temperature is no longer changing.
Reset action, however, keeps the fuel rate decreasing because the
process is off setpoint, and ceases to enter the process only when the
process has returned to setpoint. With the two control methods combined,
228 THE ROTARY CEMENT KILN

Fig. 15.14 A typical summary of proporlional, rcsct, and derivative action


combined on one chart.

system. It is also apparent that programming a controller is considerably


more complicated than manual control. Finding the correct setting is a
time-consuming task requiring preliminary process studies and process
knowledge before the correct answer can be found. One should always
keep in mind that the rewards from serious efforts made in this direction
can be large not only in clinker quality and operating efficiency but also in
the ease of operation. Without any doubt, one thing that gives the kiln
operator the most satisfaction on the job, is to have the kiln operation so
stable that only one or two small adjustments must be made during the
entire shift.
INSTRUh4ENTATION 229

15.3 KILN-CONTROLLOOPS

As previously mentioned, kiln-control concepts have become quite


elaborate and sophisticated. Automation started with simple individual
control loops that involved one input to a controller which in turn sent a
signal to the output device to turn a motor-activated valve, damper, or
drive. Technology makes it possible today to transmit and to process
hundreds of process-variable inputs simultaneously to a process computer.
This microprocessor, in turn, might be programmed and made capable of
making a control-adjustment decision for one single output based on
simultaneous analysis of ten to twenty different pieces of input data.
Single Closed-Loop Control. In single closed-loop control, a
measurement of the variable (such as temperature, pressure, etc.) to be
controlled is fed to the controller (or computer) which compares it with a
setpoint and makes any adjustment in the output if there exists an error or
deviation from this setpoint.
Some of the more common loops in this class are:

SINGLE CLOSED-LOOP CONTROL


Inpur output

Fan inlet total pressure + Fan damper position


Hood pressure + Cooler stack damper position
Undergrate pressure -+ Cooler grate drive speed
Burning-zone temperature -+ Fuel rate
Coal mill outlet temperature + Mill inlet ambient air damper
Cyclone outlet temperature + Cyclone inlet ambient damper
I.D. fan inlet temperature + Water spray rate or ambient damper
Kiln speed + Feed rate

INTERLOCK LOOPS
Inpti output

I.D. fan down + Fuel valve inoperative


Primary air fan down + Fuel valve inoperative
Gas analyzer (CO for > 2 min) + Power to electro prec. off
Coal mill outlet temp. above limit +C02tocoalmillon
230 THE ROTARY CEMENT KILN

Cascade (Computer) Control. This is the application area that is


perfectly suited for computers because in cascade control several process
variables are evaluated before a decision about a necessary change is made.
Take for example the aforementioned burning-zone temperaturejuel rate,
closed-loop control concept. In this single loop the fuel rate is adjusted
based strictly on the deviation from the setpoint of the burning-zone tem-
perature, regardless of the oxygen content in the kiln gas or the reasons for
the drop in temperature. Kiln operators would agree that this would be a
very primitive way of controlling the burning-zone temperature because
they themselves would not take such a simplistic approach when they
control the kiln manually. For example, the operator’s thought process
would follow this path:

1. Notice that burning-zone temperature is dropping


2. Questions: a)current and previous percent oxygen in exit gas?
b)current and previous kiln-torque or kiln-drive amps?
c)current and previous kiln exit-gas temperature?
d)current and previous I.D. fan speed?
e)current and previous feed loading in kiln?
f) current and previous kiln speed?
g)current and previous fuel rate?
3. Decision: If a to g = this or that, then adjust . . .

The change in burning-zone temperature is the result of a disturbance


that exists or happened earlier in one or several of the above-listed
variables. In cascade control, this mental work is being done in a split
second by a computer which evaluates these different variables, looks at
their trends, and makes the adjustment based on a program that is stored in
its memory. Advanced computer programs are capable of predicting the
change in burning-zone temperature before it actually takes place or
becomes noticeable to the operator. Coal-grinding and cooler circuits are
other areas of the kiln system where this control concept has found wide
acceptance. Complete automatic-control programs and systems are in
existence today that can take a kiln from a cold start, control it for an
indefinite length of time, and take it to a complete shutdown, without the
operator having to turn a knob. Such modem control systems usually also
incorporate graphic flow diagrams and trend displays of any circuit or
variable the control-room operator chooses. Most important of all is that
all these systems have manual control back-up capabilities to allow an
INSTRUMENTATION 231

operator to take over with manual control when the computer is down for
maintenance. Hence, regardless of how sophisticated or automated a kiln-
control system is, there is clearly a need to train control-room operators in
manual control of the kiln.
One can also discount the fear that, as the evolution of computer
control will continue, there ultimately will be no more need for kiln
operators. Perhaps it is true that the operator will have very little work to
do in the future, however, there always will be a need for this skilled
position. The operator’s job will become easier and the mental stress
factor, that was so prevalent in the “old’ times of complete manual
control, will undoubtedly be greatly reduced in the foreseeable future.
16.

Kiln Control Variables

16.1 BURNING ZONE

The burning-zone temperature is not only one of the most important


kiln control variables but also the most difficult one to monitor. Despite
the fact that burning-zone conditions in modem kilns are measured and
monitored by means of sophisticated instruments, kiln operators should
always be thoroughly trained to visually observe and evaluate the burning
zone. This capability for visual inspection is an absolute requirement
because there are times when these instruments are out of service. Visual
verification is also needed when instrument readings are questionable due to
dust interference in the burning-zone environment.
This control function is not merely a check to see whether good clinker
is being burned, but is also an evaluation of the clinker, the coating, the
flame, and the air stream. Color is important in evaluating the clinker, the
coating, and the flame. General behavior of the clinker, air stream, and
flame can be studied, as well as shape of flame and coating, and size of
clinker being discharged.
The pyrometer records the temperature of a comparatively small area. A
change in temperature registered by this instrument does not always mean
a true overall temperature change in the burning zone, and therefore does
not always indicate a need for adjustment in the kiln controls. As a matter
of fact, a change in pyrometer reading could be caused by a shifting of the
burning zone due to a change in the flame characteristics, or by dust inter-
ference entering the burning zone with the combustion air from the cooler.
With reference to the television monitor, it must be kept in mind that
the camera does not “see” the burning zone as a whole. What can be seen
on the TV screen is only a fraction of the zone. Here again, dust inter-
ference could make visibility in the burning zone poor or even impossible
232
KILN CONTROL VARIABLES

at times. Finally, either one or both instruments could fail. All these
factors emphasize the need for training a new operator thoroughly in the art
of “eye-balling” the burning zone,
Viewing Uze Kiln Interior. At this point it is important to learn how to
look properly into the burning zone. Although this seems elementary, it
should not be forgotten that the burning zone is an extremely luminous
light source. Although filtering glasses are used, the light source is so
strong that focusing the eyes into it for too long a time could cause partial
blindness. One should look no longer than one minute at a time into the
fire. If longer viewing is required, look aside for a few seconds occa-
sionally to rest the eyes. Looking steadily too long at the flame results in
the eye losing its ability to see details.
The question of what type of colored filter glass to use must be left to
each individual operator, as one person can see better with one particular
glass than with another, but the same glass may not suit someone else.
Burning with a natural gas flame usually requires a darker colored glass
than an oil or coal fire would, because of the greater luminosity of a gas
flame. As a rule, one should always use a glass that enables him to see
under and behind the flame. Once a certain glass has been chosen, the
operator should stay with this glass at all times in order that proper
judgment of the conditions in the burning zone can be made. Sometimes
an operator has the habit of overburning the clinker, and another may have
the reputation of consistently underburning the clinker. A very effective
measure to counteract these habits is to equip the “hot” operator with a
brighter glass and the “cold” operator with a darker glass to compensate for
their misjudgment of the burning-zone temperature.
Many control rooms, in modem cement plant &sign, are located far
away from the firing floor and thus demand more reliable and accurate
instrumentation to monitor the burning-zone temperature. Some of these
plants are highly successful in this endeavor but there are also plants where
this is a source of concern. Generally, it is easier to satisfy the require-
ment of recording accurate burning-zone temperature trends on kilns with
little dust interference. On the other hand, dry-process kilns, with their
commonly poor visibility, are much more difficult to monitor for buming-
zone temperature. Although desirable, absolute accuracy as to specifying
prevailing temperatures is not so important as the true temperature trend
that takes place in the kiln. In other words, as long as the instrument
registers a corresponding temperature drop when the kiln is actually cool-
ing down and an increasing temperature when the kiln heats up, the
requirement of temperature monitoring of the burning zone is fulfiiled.
234 THE ROTARY CEMENT KILN

How frequently should one look into the burning zone? There is no set
answer to this question. Naturally, on remote-controlled kilns there is no
need for the operator to pace the firing floor as frequently as on kilns that
are almost exclusively manually controlled. It is on these manually
controlled kilns where experienced operators sometimes become overcon-
fident and think that it would be perfectly safe to leave the kiln alone for
periods in excess of 40 min. This action, however, is against good bum-
ing practice. The secret of every good operator is his ability to recognize a
change in kiln condition at the time a change takes place and not later.
For this reason a good operator will never leave a kiln unchecked for too
long a time. When things are going smoothly, the kiln should be checked
every half hour, with more frequent checks if adjustments are being made.
Appearunce of CZin?w-. The quickest, although not the most accurate,
means to check the clinker for quality is by observation of the color and
the size of the clinker. Some operators have become so proficient that
they can tell fairly well how good the clinker has been burned by merely
looking at a handful of clinker. It should be pointed out, however, that
clinker size alone does not give a true indication of clinker quality as
clinker size is influenced by such factors as feed composition and kiln
speed. Generally speaking, well-burned clinker is dark, almost black in
color, and the “hotter” the clinker is burned, the larger the clinker nodules
become.
If the clinker is overburned (that is, burned hotter), the free-lime content
drops, the liter weight gets heavier, the clinker gets larger in size, more
dense (less porous), and darker in color, compared with a clinker that is not
overburned. Conversely, underbuming the clinker (burning colder) causes
the free-lime content to increase, and the liter weight to lighten. Also the
clinker gets smaller in size (dustier), more porous, and the color is
lighter-more nearly brown.
Until now the appearance of the clinker after it has been burned and
cooled has been discussed. Now consider the clinker in the burning zone.
For this a reliance on eyesight is necessary.
The kiln feed, when it approaches clinkering temperatures, undergoes
drastic changes in its physical and chemical characteristics. As soon as
liquid formation of the constituents begins, the feed becomes sluggish
because the chemical reactions tend to make the feed soft and viscous.
Formation of balls of feed (clinkers) now commences, and the material
starts to ride up higher on the rising wall of the kiln. Instead of sliding
down the wall, the feed bed starts to cascade over itself. By noting this
KILN CONTROL VARIABLES 235

cascading action in the vicinity of the flame the observer will notice an
increase in sluggishness of the feed at higher clinkering temperatures, the
feed will climb higher on the kiln wall, with more turbulent cascading.
Assuming that other variables remain constant, larger clinker balls are
formed at higher temperatures.
Now consider the color of the clinker. Any color seen in the burning
zone can be directly related to the temperature. Because different filter
glasses are used for viewing, it is difficult to assign a definite temperature
to any certain color. In very general terms, however, the kiln condition
can be estimated from the colors observed in the hottest part of the flame:

dark red cherry red orange-yellow White

cold normal hot

Any deviation of color from the orange-yellow should be investigated to


determine the cause and what adjustments need to be made to the kiln
operation.
The Feed Behind the Flame. The manner in which the operator controls
the position of the raw feed has a significant influence on operation of the
kiln, and is one of the important key indicators of operating stability.
Changes in this raw-feed posiiton are an early warning signal that buming-
zone conditions are about to change.
Special attention has to be given to the physical characteristics of the
feed behind the flame. Although in most cases it will be difficult for the
operator to see behind the flame because of the flame shape and dust inter-
ference, nevertheless, no effort should be spared in trying to see as far back
as possible, because it is in this region of the kiln where an early detection
of possible kiln upsets can be made.
The Dark Feed. When one looks into the burning zone, one will ob-
serve a sharp color change of the lower part of the feed bed under the flame
from dark to bright, as shown in Fig. 16.1. This point in the burning
zone is of great importance to the operator. The importance of this lies in
the fact that the position of this dark feed gives the kiln operator one of the
earliest indications of when the burning zone tends to warm up or cool
down. Under normal and stable conditions the position of the dark feed
remains stationary approximately one-quarter of the distance into the flame
(as shown in the figure). If it moves farther under the flame (towards the
front of the kiln), the burning zone is cooling down; if it shifts in the
direction of the kiln rear, the burning zone is warming up.
236 THEi ROTARY CEMENT KILN

Fig. 16.1 The dark feed should be about one-quarter of the way under the
flame. If the dark feed slips too far under the flame, as at A, its relative
position can sometimes be changed by adjusting the flame length, as at B.

Any change in flame characteristics will have a direct effect on the


position of the dark feed. For example, whenever the flame is shortened
the dark feed will move closer to the discharge end of the kiln. A change
in position of the dark feed therefore does not always mean that the
burning zone is either heating up or cooling down. The position of the
dark feed can move because of a change in the shape of the flame, a change
in burnability of the feed, or a change in the feed loading of the kiln. The
dark feed will move closer (under the flame) when the flame is shortened,
the feed is harder to burn, the feed loading of the burning zone increases, or
the burning zone cools down. It will move away from the flame (toward
the feed end) when the flame is made longer (and enough heat is available
to drive the dark feed back), the feed is easier to bum, the feed loading in
the burning zone decreases, or the burning zone heats up.
Any change in the position of the dark feed must be viewed in the light
of all of these influences. The operator must be able to see the dark feed
whenever he looks into the kiln, and he must regulate operation of the kiln
so as to achieve this end. A slow shifting of the dark feed in either
KILN CONTROL VARIABLES 237

direction can usually be counteracted by a small change in the fuel input


rate in order to keep the feed in its proper place.
The dark feed should never be allowed to move further under the flame
than one-half of the flame length. When it becomes necessary to counter-
act the condition in which the feed has “slipped” too far under the flame,
the operator can, if the means to do so are available, change the flame
length to restore the dark feed position to the proper relation with the
flame again. Fig. 16.1 shows how this is accomplished. Figure A depicts
a condition in which the dark feed has advanced too far under the flame. By
increasing the primary air pressure and temperature, and if possible the
secondary air temperature, the flame will be shortened as shown in Figure
B. Note that the position of the dark feed has not changed, but the relation
between feed and flame is in its proper perspective again because the flame
is shorter. This shortening of the flame results in more heat being released
in the critical area where it is most needed
Some operators take advantage of this procedure quite freely whenever
they encounter the described condition. However, one has to consider an-
other aspect caused by such a change in the flame structure. The shorter
fire results in a bushier flame that is more in contact with the coating, the
coating being thus exposed to a greater heat. For this reason a shortening
of the flame should never be carried out before the operator has assured
himself that the coating will be able to withstand the extra heat. For
example, if a hot spot on the kiln shell is already in existence in the
critical area, it would be a serious mistake to shorten the flame and release
more heat over the already weakened area.
Combuslion Air From the Cooler. Considering only the general
appearance of the air coming from the cooler under normal operating
conditions, not much attention need be given to this factor because the
appearance remains nearly constant. Under upset conditions, however, this
factor becomes very significant.
When insufficiently burned clinker is allowed to enter the cooler, the
airstream carries a large amount of fine dust particles back into the kiln
which can obstruct visibility in the burning zone. Furthermore, the sec-
ondary air temperature is most likely low in such a case, causing a change
in the flame structure, with the result that the ignition point of the fuel
moves further into the kiln. Combustion conditions are poor (cold flame)
because of the dust-entrained atmosphere in the burning zone. All these
conditons may make the burning zone appear to the eye to be “cold,” al-
though this is not necessarily the case at all times. The area of highest
238 THE ROTARY CEMENT KILN

intensity could already have approached normal clinkering temperature


again although the front is still black and cold. Whenever dust obstructs
the view in the burning zone, the operator must make a special effort to
see under and behind the flame. Experience has established the rule that,
whenever dusty conditions prevail in the burning zone, the clinkering con-
ditions under and behind the flame, rather than the color of the front of the
burning zone, govern subsequent corrective measures after an upset.
Color of the Coating. The color of the coating tells a great deal about
temperature conditions in the burning zone. Under normal operating con-
ditions, the color of the coating in the hottest area ranges between yellow
and white. When the color changes to orange or red, the zone is cooling
down; if it changes to white, the burning zone is heating up.
A large portion of heat is transferred to the feed by radiation and conduc-
tion from the kiln wall (see Chapter 2). Thus it becomes understandable
that the temperature of the coating is very important for the burning pro-
cess. In addition, the coating acts as a heat storage in the burning zone.
Having a hot coating will in many instances enable the operator to fight
out a heavy onrush of feed. On the other hand if the coating loses tem-
perature rapidly at the time a heavy load of feed enters the burning
zone,there is usually no alternative but to slow down the kiln to avoid bad
clinker.
Besides the color of the coating, the general appearance of the coating is
also important The operator should try to detect weak spots, loss of coat-
ing, or formation of rings in the burning zone. Early detection is of
utmost importance whenever a change in the coating structure takes place.
It is the operator’s duty to maintain or rebuild the coating to protect the
refractory and the kiln shell from damage by overheating.
A condition that can cause a great deal of damage to the kiln if it is not
corrected imediately is when the burning zone becomes so overheated that
the clinker starts to ball up or even worse, begins to liquefy. This con-
dition is extremely hard on the coating as the coating, becoming soft
(liquefying), starts to come off. Whenever this happens, the operator has
to disregard the production of good clinker, and concentrate on protection
of the refractory and the kiln shell. Failure to make immediate corrections
can result in red spots appearing on the kiln exterior, a sure indication that
something is wrong.
In summary, it is important to visually inspect the burning zone for
changes in the following items (any of which could signal the possibility
of change in burning-zone temperature):
KILN CONTROL VARIABLES 239

a) Clinker color
b) Clinker size
c) Cascading action of the clinker bed in vicinity of the flame
d) Feed-bed appearance behind the flame
e) Dark feed position
0 Appearance of secondary air coming from the cooler
8) Coating conditions
h) Flame shape and color
Knowing what to look for in the burning zone and being able to
recognize changes in the items listed above doesn’t necessarily indicate that
one is now ready to control the burning-zone temperature. Simply stated
and as mentioned earlier, burning a kiln is not merely a function of adding
fuel when the burning zone gets cool or conversely, reducing fuel when the
burning zone gets hot. Burning-zone temperature control must be con-
sidered with other kiln control variables before corrective action can be
taken.
Since it is difficult to measure the actual burning-zone temperature with
pyrometers, several (old and new) methods have been tried to relate other
kiln variables with possible changes in this burning-zone temperature.
The most recent and noteworthy method in this respect is the monitoring
of the NO, content in the kiln exit gases and relating this to the buming-
zone temperature changes (See Fig. 16.2).
Kiln burning is a matter of detecting any pertinent changes as early as
possible and making countering adjustments in small steps. It can best be
described as a control by anticipation, i.e., trying to establish what the
kiln conditions (burning-zone temperature) will be a few minutes later.
Waiting to make an adjustment until the full effect of a large deviation is
registered on the instrument can lead to unstable operation and is not
consistent with good burning practice. This applies not only to buming-
zone temperature but to all other main control functions.
How can an operator predict what the burning-zone temperature will be
in, e.g., 10, 20, or 30 min hence? The proven indicators of approaching
changes in burning-zone temperature are:

a) The kiln drive torque or amperage.


b) The oxygen content in the exit gases (provided the fuel rate and
I.D. fan speed remained unchanged).
c) The back-end temperature (chain outlet, kiln inlet, calcining zone)
at that time that equals the travel time of the feed from the
240 THE ROTARY CEMENT KILN

Fig. 16.2 The kiln slowdown was started in stages at 450 PM and completed
at 5:30 PM. Then the speed was gradually increased again, as the kiln warmed
up, and the kiln was back to full speed and balanced out at 12 AM.

respective thermocouple to the inlet of the burning zone.


d) The aforementioned NO, content in the exit gases.
e) The CO2 content in the exit gases.
f) The kiln-feed end draft.
These constitute the primary early-warning instrument readings that
alert an operator to upcoming changes in burning-zone temperature.

16.2 KILN EXIT-GAS ANALYSIS

Oxygen, carbon monoxide (CO), and carbon dioxide (CO2) have been
extensively discussed in Chapter 5. From these discussions it has become
apparent that the exit-gas analyzer combined with its 0, and CO recorder
KILN CONTROL VARIABLES 241

in the control room is one instrument no kiln can do without. As a


measure of safety precautions this is also one of the instruments that
irrevocably MUST be in operation at all times while the kiln is operating,
i.e., when fuel is fired into the kiln.
This gas analyzer, because it samples “dirty” kiln gases and takes the
sample at a location where high temperatures prevail, has a tendency to
malfunction frequently unless almost daily preventive maintenance is
carried out on this unit. The location, where the sample probe is installed,
is also a key point to consider as false air inleakage could distort the true
contents of 02 and CO, in the exit gases. Kiln operators must be on the
look-out for such malfunctions and notify the foreman or instrument shop
whenever a malfunction occurs or is suspected. Operating a kiln without
the gas analyzer functioning properly is a dangerous thing to do and can
lead to catastrophic accidents. This is one instrument that doesn’t allow
room for compromises; it simply has to work and must be repaired as
quickly as possible when it doesn’t. These are strong words but considered
worth mentioning for the simple reason that many accidents have happened
in the past in some cement plants because the gas analyzer wasn’t working
properly.
For example, a few years ago a kiln operator noticed an irregularity in
the gas-analyzer recording wherein more than 2% carbon monoxide (CO)
was indicated but, at the same time the oxygen content showed peak levels
of 5%. In this example, the operator made the wrong assumption,
specifically he believed the oxygen (02) to be correct and the CO recording
to be in error. Result: A fire in the kiln back end. Lesson: Anytime that
the presence of carbon monoxide (CO) is indicated, reduce the fuel rate
immediately and ask questions later. In such situations it helps to re-
member the triangular relationship for combustion discussed in Chapter 5.
In the above example fuel was present (i.e., CO which is unburned car-
bon), air for combustion became available from possible inleakage of
ambient air into the system at the kiln back end, and the only missing link
for a fire or explosion at the feed end was heat for ignition. The third
component for this accident to occur was most likely provided by the high
temperature of the exit gas or the electrostatic precipitator which by itself
is an igniter in the true sense. To safeguard against such accidents, most
electrostatic precipitators are interlocked with the fuel rate. In other words,
the precipitator is automatically deenergized whenever CO is present for
more than 1 min in the exit gas.
Earlier, it was mentioned that both too much (more than 3.5%) and too
242 THE ROTARY CEMENT KILN

little (less than 0.7%) oxygen in the kiln exit gases represents inefficient
operation and that l&1.5% 02 is a level that results in optimum operat-
ing conditions. This applies to all kilns with the exception of a
precalciner kiln that has no tertiary air supply to the flash calciner. In
these types of precalciner kilns, all the combustion air requirements for the
flash calciner are supplied from the kiln, hence higher oxygen levels are
required at the kiln exit to secure complete combustion of the fuel given to
the flash furnace. However, the same principle would apply at the
preheater cyclone outlet, i.e., for the gases after the flash furnace. Here
too, the oxygen content should not be less than 0.7%, not more than
3.5%, and ideally between l&1.5% during normal kiln operations. In
plants where low-grade fuels with large variations in heat value are used for
firing in the flash furnace, it is advisable to hold the percent oxygen at a
higher level than the indicated range to secure sufficient air availability
when surges of higher heat-value fuels occur. Since precalciner kilns
operate with two combustion-process locations, it is advisable to equip
these kilns with two gas analyzers, one at the kiln exit and the other at the
flash calciner outlet. In this manner, both combustion processes can be
independently controlled from each other. However, on precalciners that
are equipped with only one analyzer at the preheater tower exit, an operator
might have considerable difficulty in relating changes in 02 or CO to
either one of the two burning locations.
The question arises: “What is to be done when the kiln operates steadily
but continuously with oxygen contents higher than say 3% or lower than
0.7%?” In such instances, because of prevailing stable conditions, no
drastic changes should be made as such moves could upset the delicate
balance of the kiln. Rather, it is advisable to fine-tune the controls in
small steps, allowing ample time of about 1 h between each to make sure
this balance and stability is not disturbed Fine tuning in this manner may
take up to 8 or 16 h and is referred to as optimizing the kiln operation. As
with any other kiln control variable an operator follows the following
basic steps in mattters of exit-gas control:

1. Secure the safety of the equipment and personnel (sufficient O2


available and no CO showing)
2. Stabilize the kiln (02 to level off and stay within a certain narrow
rws)
3. Optimize (fine-tune) the O2 in small steps to bring the oxygen
content into the range of l.OO-1.5% to obtain optimum kiln
efficiency.
KILN CONTROL VARIABLES 243

The majority of kilns are equipped with analyzers that test only for
contents of oxygen and carbon monoxide which are considered the two key
variables in combustion control. Some kilns also have separate carbon
dioxide analyzers and monitors (see Chapter 5) but these are not considered
as absolute requirement, rather they are excellent tools to forewarn the
operator of upcoming changes and are helpful in optimizing the kiln
operation.
In recent years, a great deal of work has been done in some plants in the
utilization of recordings of NO, in the kiln exit gases for determining
changes in burning-zone temperature. Originally these NO, analyzers were
installed and used for the purpose of emission control to satisfy
environmental specifications. Engineers, being familiar with the process
of NO, formation in a rotary-kiln flame, have recognized the relationship
between the NO, in the exit gas and the burning-zone temperature.
Different NO, concentrations are found in the exit gas when the com-
bustion temperature (flame temperature) in the burning zone changes.
Higher flame temperatures result in higher NO,. Furthermore, when the
kiln excess air (02) increases, the NO, also increases. It must be
mentioned that NO, cannot be related to the actual prevailing temperatures
in the burning zone but it is an excellent and rapid indicator of changes
that have taken place there. In many respects the above reactions to flame
temperature and excess air in the kiln lead to contradictory objectives in
kiln operation and have been the subject of much controversy. Naturally,
in matters of environmental control, one strives to obtain a NO, content
as low as possible. For an efficient kiln operation, however, one tries to
operate the kiln with as short a burning zone as possible (short flames,
high flame temperatures) which in turn means high NO, contents. This
problem is especially acute on natural gas-fired kilns because they usually
operate with higher NO, contents than coal- or oil-fired kilns.
As a general rule, kiln exit-gas components are affected as follows,
assuming all other variables remain constant.
Strip chart recordings of oxygen often display large variations in range
even when the kiln operation is stable. This is referred too as process
“noise” and is a normal occurrence on many kilns. However, when this so-
called “noise” exceeds a range off 0.5% in short-time intervals, this could
be an indication of irregular fuel or air-flow rates taking place within the
system and should be investigated. Likewise, since process “noise” is
common to the oxygen analyzer, the instrument should be checked when-
ever the oxygen chart traces a straight line for a long period of time.
Most oxygen-recording charts register this variable within the range of
244 TKE ROTARY CEMENT KILN

02 CO Co, NO,

Fuel increases down UP down down


decreases UP down UP UP
I.D. fan increases UP down UP UP
decreases down UP down down
Flame temperature increases = E: = UP
decreases = = = down
Feed rate increases down = up =
(calcination) decreases UP = down =

O-5%. It is important to remember the following: When the recording is


pegged out at the maximum for a long period of time (i.e., draws a straight
line at the top), one doesn’t know if the oxygen is at 5.1 or at 15%. In
short, an operator under these circumstances wouldn’t know what the com-
bustion conditions are. It is therefore good burning practice to always set
the fuel rate and/or I.D. fan speed so that the oxygen recording is always at
least below the 5% mark whenever fuel is being fired into the kiln. This
recommendation also applies and is of equal importance during times of
kiln starts and warm-ups. As a matter of fact, as soon as a main fire is lit
in the kiln, it is the operator’s first duty thereafter to stabilize the oxygen
content at a level that is at least below the 5% oxygen mark.

16.3.FUEL-RATECONTROL

Before any main fire (except the kiln warm-up torch) can be lit in the
kiln, one has to ascertain that the following requirements are met:

a) Sufficient air is present to achieve complete combustion.


b) The gas analyzer is operational and functioning properly.
c) SuffZznt heat is present to readily ignite the fuel.
d) The I.D. fan is running and the kiln draft is properly regulated to
prevent an abrupt, delayed, explosive ignition of the flame.
e) The firing floor is cleared of any unauthorized personnel.
I) The primary-air fan is running and its flow rate properly set.
KILN CONTROL VARIABLES 245

Once the fuel has ignited and a proper flame obtained, the operator must
immediately check the gas-analyzer recording to ascertain that no
combustibles (CO) are showing after a time delay of = 30 s. After this,
necessary adjustments (fuel or ID. fan) to bring the oxygen (02) content
below the 5% mark must be made. These am standard and elementary
steps that apply to any kiln when lighting the main fire. Another rule that
is practiced in most plants for safety reasons is to cut off the fuel com-
pletely whenever the flame is not lit after 30 s. The kiln is then purged
for at least 5 min to rid the kiln of any combustible gases before another
attempt is ma& to reignite the fire.
The operator should never attempt to control burning-zone temperature
by merely increasing or decreasing the fuel rate, nor should he, for
example, increase the I.D. fan speed to permit raising the fuel rate because
the exit gas is deficient in oxygen. To operate in this manner is sure to
lead to trouble, because no consideration is being given to the temperature
at the back end of the kiln, where the feed is being prepared for calcination.
Increasing the I.D. fan speed in order to be able to add more fuel (in case of
low oxygen) will cause the back-end temperature to rise. This however is
the wrong thing to do, because the back-end temperature plays a vital role
in the operation of the kiln, and cannot be allowed to freely float up and
down.
Whenever the operator makes a change in either the I.D. fan speed, the
fuel rate, or both, he has to anticipate the possible reactions caused by the
change, remembering that an increase in fuel rate results in higher back-end
temperature, low percentage of oxygen in exit gas, and higher buming-
zone temperature, and an increase in I.D. fan speed will result in higher
back-end temperature, more oxygen in the exit gas, and lower burning-zone
temperature. From this it becomes obvious that the fuel rate alone does
not govern the burning-zone temperature. Without changing the fuel rate,
a change in the burning-zone temperture can be caused merely by altering
the kiln draft with I.D. fan speed. Nevertheless it should be emphasized
that a fuel-rate adjustment will give the fastest reaction whenever a change
in burning-zone temperature is required
The first rule in fuel-rate control is: Always check the oxygen and CO-
analyzer recording before and after any fuel-rate adjustment is made. It will
not take long for a new kiln operator to learn that the burning zone will
react slowly to any fuel-rate adjustment. There is an inherent time delay
until a noticeable change in temperature takes place. Since these time
delays can be as much as 10 min, particularly after an upset operating con-
dition, it is quite common among new operators to either over-fuel or
246 THE ROTARY CEMENT KILN

under-fuel the kiln for a prolonged length of time. The end result of this
action is a burning zone that alternates from “cold” to excessively hot, a
condition generally referred to as a cycling kiln. Because of this time
delay, operators must become skillful in anticipating these swings of ex-
treme temperatures and make their fuel adjustment before the actual change
in temperature takes place. For proper timing of these adjustments the
operator checks other instruments which could give him a positive sign
that a turnaround in temperature conditions in the burning zone is
imminent.
These so-called early warning signals are:

Burning zone
Oxygen(%)-without fuel increases: heating up
or fan adjustment decreases: cooling down
Kiln-drive torque increases heating up
decreases cooling down
NO, WI increases: heating up
decreases: cooling down
co2 PO) increases: cooling down
decreases: heating up

I Back-end temperature* was higher than optimum:


was lower than optimum:
heating
cooling
up
down

* Refers to the temperature conditions that prevailed at a time that is equal to the travel
time of the feed through the kiln, e.g., if travel time = 1.5 h, the operator looks at
the back-end temperature that prevailed 1.5 h before.

The time lag between burning-zone temperature and a fuel-rate adjust-


ment has been previously mentioned. With direct-fired coal systems this
lag time is even more pronounced since it takes several minutes between
the time the coal rate to the mill is adjusted and the change to be noticed at
the burner tip.
Much work has been done with automatic fuel-rate control. On pre-
calciner kilns, the fuel rate to the flash calciner is usually automatically
controlled in a closed loop that uses the fourth-stage, preheater cyclone exit-
gas temperature as the input variable. Simply stated, when this
temperature is below the setpoint, more fuel is given to the flash calciner
and, conversely less fuel when the temperature is above this setpoint.
KILN CONTROL VARIABLES 241

Constraints in this loop are the same as with the fuel-rate adjustments at
the rotary-kiln firing end, namely the oxygen must conform to certain
predetermined levels and no CO is allowed to show. Since a flash furnace
is much smaller than a rotary kiln, the controller must be tuned so that
there will be a fast response by the coal feeder whenever a change in this
fourth-stage, cyclone outlet temperature takes place.
Conventional fuel-rate control on wet, dry, and preheater kilns can be
manually or automatically controlled. When automatic control prevails,
the control concept should preferably consider as input variables at least
the burning-zone temperature, the back-end temperature, and the percent
oxygen in the kiln exit gases. A simple control logic of adjusting the fuel
rate strictly and solely based on the burning-zone temperature alone seldom
yields satisfactorily stable and efficient kiln operations.
During times of severe kiln upsets, i.e., when underburned fine clinker
has entered the cooler and “blackened” out the burning zone, an operator
often tends to make the mistake of overfueling the kiln. This is a natural
tendency of all operators because a lot of heat is needed to bring the bum-
ing-zone temperature back to normal. Since the kiln speed is drastically
reduced in such instances and visibility is extremely poor, they tend to
leave the fuel rate at high levels for too long a period of time waiting for
the burning zone to clear. However the kiln could already be in an over-
heated condition. Things to consider during these conditions are:
a) Operating the kiln with no or very little oxygen contents (even
when no combustibles are showing) often does not produce the
desired heating of the burning zone. Reason: The dust in the kiln
and low oxygen combined together produce lower flame
temperatures. Solution: Try a little less fuel to operate at a
slightly higher oxygen content of, e.g., 0.549%.
b) Once a heavy onrush of material has passed through the burning
zone, there invariably is a lighter load behind it that causes a rapid
increase in temperature when it arrives in the burning zone. This
rapid temperature increase can usually not be detected early
enough because of the dusty material in the cooler. Solution:
Watch for a definite sign that the load is becoming lighter
(oxygen increases, NO, increases, kiln amp increases, cooler
undergrate pressure decreases) and start to increase the kiln speed
at that time. Depending on the severity of this light load it
might become necessary to start reducing fuel rate at this time
also (the effect on back-end temperature should not be forgotten)
248 THE ROTARY CEMENT KILN

to prevent the burning zone from becoming too hot. In short,


this is a time where anticipation of upcoming changes becomes
of primary importance.

16.4 KILN-SPEED CONTROL

A kiln can never be expected to operate in a stable condition for an


indefinite length of time because changes do take place in the kiln regard-
less of how stabilized the operation may appear. Sooner or later an adjust-
ment in kiln speed has to be made if the kiln is to continue to produce a
good clinker. The commonest action the operator must take in case of an
apparent upset is the so-called “slow-down” procedure, as described in the
following paragraphs, and shown graphically in Fig. 16.2.
After the kiln has been operating for a long period of time under
balanced conditions the operator detects a heavy onrush of feed directly
behind the flame. Whenever this happens, the operator must determine
whether he will be able to hold the heavy feed load with an increase in the
fuel rate alone, or are conditions such that the dark feed could pass too far
under the flame even with the fuel rate increase? It is important that this
decision be made at the earliest time possible. Oxygen percentage in the
exit gas, back-end temperature, and conditions in the cooler are the deciding
factors to be considered.
Assume that all indications are that it will not be possible to maintain
the same kiln speed, and therefore speed will have to be reduced. It then
becomes necessary to determine how much the kiln can be slowed.
Depending on the magnitude of the “push,” 5-10 rph less will be suffi-
cient in some instances. In other cases the onrush is so heavy that the
kiln has to be slowed down to minimum speed. Only experience will tell
the operator how much the kiln has to be slowed. One rule, however,
applies at all times: Never allow raw, unburned feed to enter the cooler,
even if this means that the kiln has to be stopped and reheated on quarter
turns or on auxiliary drive.
After the kiln has been slowed, there are, of course, a number of
changes that take place, and the operator must be alert to watch for these
changes and take appropriate corrective action. First, the back-end tem-
perature will start to increase. This is undesirable because this temperature
should be held within f 20 degrees. Reducing the ID. fan speed will aid
in holding this temperature. Because of the decrease in the kiln draft
KILN CONTROL VARIABLES 249

resulting from the I.D. fan-speed reduction, there will be insufficient


oxygen available for complete combustion. This is corrected by reducing
the fuel rate until the exit-gas analyzer again indicates from 0.448%
oxygen in the gas.
Because of the slower kiln speed, less material will enter the cooler,
thus both secondary air temperature and undergrate pressure will decrease.
Therefore, reduce the bedgrate speed in the cooler to maintain more or less
the same undergrate pressure. This will be difficult to do. On very slow
kiln speed it is almost impossible to obtain the same undergrate pressure
as on full speed. Under slow-speed operation, do not attempt to hold the
secondary air temperature on the same level as on full kiln speed, espe-
cially when the clinker is slightly underburned, because of the danger of
the clinker not being sufficiently cooled when it leaves the cooler.
While the kiln is on slow speed it is necessary to decide how scan the
speed can be raised again. First of all, never increase the kiln speed before
there is a definite sign of the burning zone warming up, or there is a defi-
nite indication that the burning zone will warm up within a few minutes
because of a lightening of the load. Do not hold the kiln on very low
speed until the burning zone has reached normal clinkering temperature;
start increasing the speed as soon as there is a sign of warming up.
Signals and instruments to watch for these early warnings are discussed in
Fig. 16.3. The more the kiln speed is increased, the smaller the speed
increments should be and the longer the time interval between each speed
increase. The way in which the kiln speed is increased is probably the
most important factor in getting the kiln into stable conditions again.
Consider the three examples shown in Fig. 16.3.
Example 1: This shows an ideal execution of a speed-up procedure.
The operator started to increase the kiln speed in large steps of 5 rph a t a
time until he reached kiln speed 50. From then on he extended the time
between each speed increase and carried out the increases in smaller steps as
the speed was increased. These were at the rate of 3 rph from 50 to 62,
2 rph from 62 to 66, and 1 rph from 66 to full speed of 70 rph.
Example 2: In this example the operator used the wrong judgment.
He started to raise the kiln speed too fast before the burning zone was ready
for it. Very soon he had to backup on the kiln speed and had to start re-
heating the burning zone all over again.
Example 3: Here the opposite condition occurred The operator
waited too long on slow kiln speed. Suddenly the burning zone started to
gain heat very rapidly, forcing him to raise the kiln speed in large incre-
250 THE ROTARY CEMENT KILN

TIME-

Fig. 16.3 Kiln speedup, for any reason, must be done in easy stages, as
shown in Example 1, commencing with large steps close together, then
extending the time between steps as the steps become smaller.

ments and in short time intervals. For a short time full kiln speed was
able to be maintained, but very soon the kiln became overloaded and the
operator had to slow the kiln down once again.
Remember that every time the kiln speed is increased, the back-end
temperature drops, requiring more fuel to burn with each speed increase.
This requires an adjustment in the I.D. fan speed as well as in the fuel rate.
Both are increased in a manner similar to the kiln-speed increase in
Example 1 (Fig. 16.3).
The ideal example given in Example 1 by no means suggests that the
kiln will operate in a stable manner again for a long time on full speed. In
time intervals of 2-3 h another push, each time less severe, will be en-
countered. Each time the slow-down period will be shorter and each time
the chances are better that the kiln can be held at the normal operating
speed. Sometimes one slow-down will be sufficient, at other times two or
three such slow-down sequences have to be undertaken until the kiln again
is stable. The chances of avoiding a cycle in the kiln are minimized when
the operator executes a slow-down sequence according to the ideal example
given. The aforementioned discussion applies primarily to long-wet and
dry kilns. Preheater and precalciner kilns react differently.
Kiln Rollback. The feed bed in a rotary kiln occupies up to 10% of the
cross-sectional area of the kiln. While the kiln is rotating, the center of
gravity of this material is displaced to one side (on the rising side) of the 6
KILN CONTROL VARIABLES 251

o’clock position (bottom center of the cross section), the torque resulting
from this displacement being opposed by the driving torque applied to the
kiln. When power to the drive motors is cut off for any reason, the off-
center position of the load causes the kiln to rotate in the reverse direction
until the feed bed comes to rest at the 6 o’clock position. The reverse
rotation of rollback can reach a maximum speed that exceeds the normal
running speed of the kiln, thereby damaging the drive gear.
Another problem associated with rollback is that of restarting a kiln
that has stopped while loaded. If left to rotate freely after a stop, the kiln
will come to rest with the load of the feed bed at or near the 6 o’clock
position. In the past the relatively small rotary kilns could be started from
this position without much difficulty, as the kiln drives then in use were
capable of overcoming the initial torque to set the kiln in motion. With
the advent of the 500-foot or longer giants, kiln manufacturers had to find
a way to overcome static fricton and inertia in or&r to accomplish initial
acceleration with a minimum of strain on the drive gear and motor. This
resulted in the concept of kiln rollback control. In this method of control,
instead of permitting the kiln to reverse rotation, or roll back, when the
drive motor is stopped, a brake is applied to the drive unit to stop the kiln
at the instant the motor circuit is broken with the feed-bed position in an
inclined angle to the rising side. Thus the gravitational force of the kiln
feed bed acting downard is frozen by the brake so that it can be used advan-
tageously when the kiln is started again.
When the operator presses the kiln-drive start button, the following se-
quence of actions takes place automatically; first the brakes are released
permitting the kiln to roll backwards (Fig. 16.4) with the feed bed past the
low point of the kiln circle. Next the feed bed, now being on the down
side of the kiln, causes the kiln to start to roll, or rotate, forward in the
normal direction of rotation, at which time the kiln drive motor is acti-
vated to keep the kiln revolving in the forward direction.
This system has proved its usefulness and has contributed significantly
to the present long life of drive gears and motors of large rotary kilns.
The most commonly asked question is: “What is the optimum kiln
speed for efficient and stable kim operation?’ There is no clear-cut answer
because each kiln is different in its requirement. The kiln slope and the
resultant feed velocity in the kiln is one factor that must be considered.
Another important factor is the feed loading of the kiln, i.e., the output
rate that can be achieved at a given kiln speed without sacrificing operating
stability. Typically, specific volume loading of kilns varies from a low of
252 THE ROTARY CEMENT KILN

Fig. 16.4 Kiln rollback is used to advantage when starting a kiln under load
When the drive motor is stopped, brakes are applied to hold the feed bed in
Position A. The arrow indicates the normal direction of rotation. When the
drive-motor start button is pressed to start a loaded kiln, the brake is
automatically released, permitting the kiln to roll back to Position B. ‘Ihe kiln
now starts to roll forward as shown in Position C, at which time the motor is
activated to continue rotating the kiln.

approximately 5% to a high of 10%.


Dry and wet kilns with steep slopes should be operated at slower speeds
to prevent large onrushes of feed waves into the burning zone (and into the
cooler) during upset conditions. Such kilns usually operate normally in
the range of 50-60 rph. Other less-sloped dry- and wet-process kilns
generally run at speeds of 60-72 rph whereas preheater and precalciner
kilns operate at much higher speeds. Kilns that consistently and periodi-
cally exhibit difficult burning conditions in the burning zone as a result of
incomplete feed calcination sometimes exhibit better operating stability
when the kiln-speed target is lowered and the kiln is concurrently operated
at a higher percent specific loading. This, however, is in direct conflict
with the “old” heat-transfer law for rotary kilns which states that
more efficient and favorable heat-transfer conditions exist when the kiln is
operated at high kiln speed and low percent feed loading. Hence, instead of
lowering the kiln-speed target and operating at a deeper bed depth to over-
come frequent operating instability, it might make more sense to equip
such kilns with more heat exchangers in the calcining zone to improve the
overall heat transfer behind the burning zone. Regardless of the kiln-speed
target (top speed under normal operation) it is good practice to set this
target at a speed that is approximately 5-15 rph lower than the maximum
permissible. The kiln must be allowed some “elbow room” at the top to
make it possible to further increase the kiln speed in an emergency when
the burning zone has become dangerously overheated.
17.

Fuel Systems

17.1 FUEL HANDLING AND COAL GRINDING

Cement kilns are usually fired with oil, natural gas, or coal. Gas firing
requires no fuel preparation; this type of fuel is used directly as it is being
delivered to the plant by the gas pipe line. In oil firing, the oil has to be
preheated to a given temperature that produces the desired viscosity for
proper atomization of the fuel at the burner tip. This is usually accom-
plished in heat exchangers that predominantly employ steam as the heating
medium. Automatic controls must maintain the oil temperature within a
very narrow range of + 5 C (k 9 F) which is considered essential for
uniform firing conditions and flame stability.
With gas firing, the operator’s duty is relatively simple in matters of
fuel-flow monitoring. The operator must watch out for the possibility of
sharp pressure changes in the gas supply line that would indicate some
problems along the supply line. Gas firing of the flash furnace in a pre-
calciner kiln, however, has been shown to present some problems when
compared to the other two fuels. Here, combustion conditions have to be
very closely monitored to ensure that all the gas introduced into this
auxiliary firing unit is completely burned within its confines. When this
prerequisite is not fulfilled, there exists the danger that some of the un-
burned natural gas could escape into the upper stages of the preheater cy-
clones, undergoing combustion there and ultimately causing elevated
temperatures of the gases leaving the preheater tower.
With oil firing attention must be paid to the oil preheat temperature as
any significant change could lead to a change in flame characteristics.
Furthermore, close attention must be given to the pressure indicators in
the heat exchangers because plugged filters could lead to an interruption of
253
254 THE ROTARY CEMENT KILN

the oil flow to the burner. Steam pressure must also be frequently
monitored to make sure sufficient steam is available to preheat the oil
properly.
Control of the coal-grinding plant and coal conveyance to the burner is
much more demanding than when either one of the above-mentioned liquid
or gaseous fuels is used. An operator, for safety reasons, should not be
allowed to exercise control over the coal handling and grinding plant unless
that operator is completely familiar with all safe operating procedures for
the system. Coal firing has its advantages, namely, a coal fire is more
luminous and delivers better heat transfer by radiation from the flame to
the feed. In short, there is less time lag between the moment a fuel adjust-
ment has been made and the time the clinker bed reacts to this change.
However, this advantage is neutralized in direct-fired kilns because of the
time delay between the moment coal-rate adjustments are made and the coal
has been ground and insufflated into the kiln. Operators that have had the
opportunity to fire different kinds of fuel tend to agree that the burning-
zone temperature can be more rapidly adjusted with a coal fire than with
any other type of fuel.
An entire book could be written on the subject of coal handling, grind-
ing, and firing. For the benefit of the operator, the subjects that are of
paramount importance to him, namely his safety and the safety of the
equipment, are discussed here. An operator can become comfortable with
coal firing provided that a respect for its limitations and inherent safety
requirements is developed. The operator must know the system’s idiosyn-
crasies and immediately be able to recognize a potentially dangerous
condition when it develops. Pulverized coal, mixed with excessive air and
exposed to excessive temperatures, represents a potentially explosive
mixture. High-volatile coal, even in the unground state, can also undergo
spontaneous ignition while in storage. Prerequisites for successful and
safe grinding and handling operations for coal are:

a) well-established, safe standard-operating procedures,


b) close and continuous attention to the instrumentation by com-
petent operators, and
C) regular preventative maintenance for all system components.

Even the best-&signed and maintained systems will sooner or later


present the operator with an unusual condition that demands immediate and
corrective action on his behalf. An operator must be on the lookout for
FUEL SYSTEMS 255

the following potentially dangerous conditions each of which could lead to


a fire or explosion:
l Fresh smoldering coal (i.e., coal that has started spontaneous
ignition) being fed to the coal mill
l Metal pieces, rags, and other materials in the coal being fed to the
mill causing mill, bin, or feeder plug-ups and metal fragments
not being ejected from the mill (sparks could ignite the coal)
l Mill outlet temperature too high or showing a rapid rise

l Fresh unground coal too wet (> 15% moisture) that could lead to
any one of the following conditions:
a ) insufficient drying in the mill,
b ) accumulations in the mill, bins, and/or coal pipe,
c ) excessively high mill inlet-temperature demands, and
d ) too low a mill outlet temperature
l Insufficient air velocity in the mill and coal pipes causing
settlement and accumulations of coal or possible entrance (back
flashing) of hot kiln gases into the burner pipe
l On semidirect or indirect-fired kilns: worn vanes on rotary feeder
or malfunctioning air locks-allowing either coal to seep into the
primary air pipe or hot air to enter the coal bin during shutdowns
0 Leaks in the coal pipe
l Poor housecleaning with large accumulations of coal near the coal
handling, storage, or grinding system and on the firing floor
l Entire coal system or primary air fan has shut down under load
due to a power failure, i.e., coal is present in the mill and the
burner pipe (IMPORTANT; DO NOT OPEN DOORS OF
SYSTEM FOR MANUAL CLEANOUT OF ACCUMULA-
TIONS BEFORE THE SYSTEM HAS COOLED DOWN TO
AMBIENT TEMPERATURE)
l Ground coal insufficiently dried causing settlement, plugging, and
accumulations (coal pockets) in bins, dust collectors, and coal
Pipes
l Entire coal system has not been properly cleaned (air swept) when

the kiln was stopped for a prolonged shutdown


l Operating the system with faulty instrumentation, improper
damper control, and lack of proper fire-extinguishing and
explosion-prevention equipment or devices
l Smoking by employees or welding by maintenance crew near the
system while it is in operation
256 THE ROTARY CEMENT KILN

An operator should discuss all of these conditions with his supervisor


and must know what corrective action should be taken if any of these con-
ditions arise. These procedures must be clearly and fully memorized be-
cause there will not be sufficient time available to look these up in the
standard operating procedures when such emergencies develop. I
When coal is received at the plant, it has to be dried as most coals
contain appreciable amounts of moisture. This drying is usually done in
the coal mill itself using hot excess cooler air. This hot cooler air is
tempered with cool ambient air a short distance past the cooler to prevent:

a) overheating of the dust cyclone and air pipe ahead of the coal
mill, and
b) excessively high temperatures from entering the coal mill when
overheated conditions prevail in the cooler itself.

Other plants use the inert preheater exit gases for drying in the coal mill.
Since these gases usually contain less than 5% oxygen, they represent a
safety advantage when compared to the use of cooler excess air.

IMPORTANT:

a) Know what the maximum allowable temperature is for the coal


mill inlet temperature.
b) Do not exceed this temperature under any conditions.
c) Make sure that this tempering damper is not fully open during
normal operating conditions to allow for further admittance of
cold air when the need arises.

A second tempering damper is usually installed just ahead of the coal


mill that is the primary control for the coal mill o&et temperature. Cold
ambient air is drawn into the system to hold this coal mill outlet
temperature within a narrow range regardless of changes in the moisture
content of the unground coal. Typical setpoints for coal mill outlet tern-
peratures are:

for low-volatile coal: 80-90 C (175-195 F)


for high-volatile coal: 68-80 D (155-175 F)
FUEL SYSTEMS 257

When this coal mill outlet temperature exceeds 93 C (200 F) the danger of
premature ignition of the coal in the mill exists.
Grinding of the coal is done predominantly in roller mills although
there are still several plants that use ball mills for this purpose. Both
these types of mills rely on air-sweeping action to evacuate the coal from
the mill. In semidirect- and indirect-fired systems the coal/air mixture is
blown into a holding bin whereas in direct-fired systems the coal is directly
insufflated into the burning zone. IMPORTANT: The following fun-
damental conditions must always prevail whenever any of these systems
are in operation:

a) Sufficient air has to pass through the mill to properly evacuate


the ground coal from the mill. This can be monitored by the draft
at the mill inlet and the mill outlet. Both these drafts must not
be allowed to drop below a given preestablished value. Low-
pressure differentials between these two measuring points are an
indication of mill plug up or lack of sufficient air for sweeping
the coal from the mill. An operator should know exactly what
the critical values for coal mill drafts are for the particular system
under his control. The same principle applies to dust collectors
incorporated into an indirect-firing system
b) To keep the coal particles in suspension while being transported
from the mill to the bin or burner, a minimum velocity of air
passing through the coal pipes must exist at all times. As a rule
of thumb it is advisable ll~t to let this velocity drop below 30
m/s (6000 ft/min). Since there is a direct relationship between
the mill-exit draft and this velocity, an operator monitors this
draft (on direct-fired kilns this is the primary air pressure) and
must know what the minimum permissible pressure is. The
operator will then have to make sure to never operate below this
critical pressure whenever coal is being ground in the mill.
C)‘Since one kilogram (or pound) coal requires approximately
8.5-10.5 kg (or lb) air for combustion, the fuel/air mixture
emanating from the coal mill must be kept well below this
critical ratio to prevent premature ignition. Most coal mills
operate at air/fuel ratios that are less than 2.0 kg (lb) per kg (lb)
of coal. However, direct-fired kilns, with their high percent of
primary-air requirements, usually operate at higher ratios and
therefore need closer attention.
258 THE ROTARY CEMENT KILN

d) Combustion air (primary, secondary, or tertiary air) should be


high enough to ensure continuous ignition of the fuel or light-up
torches must be lit until the fuel can maintain self-ignition.
All coal-handling and grinding facilities should be equipped with
automatic shut-down interlocks. Some of the more common interlocking
systems are:

Operating Condition Automatic Action Taken


1. I.D. fan stopped All fuel off
2. Combustibles for longer All fuel off and precipitator
thanlmin deenergized
3. Gas temperature too high in Flash furnace shutdown
precalciner
4. Primary air fan stopped Fuel off
5. Temperature too high at mill Coal mill shutdown and activation
outlet of CO2 extinguishers
6. Flash furnace temperature Fuel in flash furnace off
too low
7. Cooler exhaust fan stopped Fuel off
8. Pulverized coal bin shows Activate inert gas (CO2, N2) and
rapid temperature rise fire suppressants

The optimum coal fineness must be established for each kiln and depends
on such factors as type of coal fired and type of flame required. Sub-
bituminous and high-volatile bituminous coal are usually ground coarser
than low-volatile coal or coke. As a general guideline, the following
criteria can be used whenever the coal character&&s change:

If Then
Volatilles increase (%) Grind coarser
Volatiles decrease (%) Grind finer
Ash increases (%) Grind finer
Ash decreases (%) Grind coarser
Ash rings form at kiln outlet Grind finer (avoid oversized particles)
Plume too long (late ignition) Grind finer and/or increase secondary
air temperature
Plume too short (early ignition) Grind coarser and/or decrease primary
air temperature
FUEL SYSTEMS 259

Another factor to consider is the so-called ash-softening temperature as


this could be a primary cause for troublesome ring formations at the kiln
discharge. It is generally believed that coals with a low ash-softening tem-
perature and/or high contents of iron in the ash are prime contributors to
such ring formations. Here, too, it is recommended that finer grinding be
employed to combat such rings.

17.2 FUE!, BURNERS AND FLAMES

It is difficult to control the shape of a coal flame during the course of an


operation unless one of the modem, sophisticated adjustable burners is
used. A change in primary air pressure is about the only real adjustment
an operator can make while the kiln is in operation. Although position
and the design of the burner as well as hood draft and secondary air
temperature are prime influencing factors for flame shape, these variables
can not be easily changed by the operator.
Very complex adjustable and/or combination burners are successfully
used for oil, gas, or coal firing of cement kilns. These are excellent tools
for flame and coating control but are often misused. Many times operators
have opposing views about appropriate and desirable flame shapes and thus
make too many adjustments too frequently. This is especially true with
burner designs that are equipped with adjustable inserts to promote
turbulence and mixing of air with the fuel. Furthermore, any coal burner
that contains such inserts lends itself to potential coal-pocket formation
during unusual operating conditions. Because these adjustable burners are
so efficient in flame-adjustment capabilities, any error in adjustment
setting by the operator can directly lead to thermal abuse of the refractory
and kiln components near the kiln discharge area. This has compelled
many plants to use a straight burner pipe with no inserts for natural gas or
coal firing. However, the advantages in low costs and prevention of coal-
pocket formation is outweighed by the disadvantages and limitations of
these simplified burners when used on modem preheater and precalciner
kilns. Straight burner pipes are designed for normal “full-speed” produc-
tion operation and usually do not deliver satisfactory flame shapes and
characteristics during start-up and kiln slow-down periods. During these
times the flame is often too long and the ignition of the fuel erratic
causing dislocations in the position of the burning zone. Long, lazy
260 THE ROTARY CEMENT KILN

flames can also create excessive coating and ring formations. Operators are
usually capable of coping with this problem by using the so-called
auxiliary start-up torch as an aid in ignition which by itself is expensive
and requires frequent watching. Short-term dislocations of burning-zone
positions in wet- and dry-process kilns are not so critical but can become
disastrous on preheater and precalciner kilns. On these modem kilns the
burning-zone position can not be allowed to fluctuate as freely because
these kilns operate at much higher speeds and are considerably shorter in
length. Another disadvantage of the straight burner pipe is the need for
redesigning the burner geometry (for proper tip velocity) whenever there is
a major modification made on the kiln that would produce a significant
change in the specific fuel requirements.
All fuel burners are exposed to severe wear and heat conditions inside
the kiln and should therefore be frequently inspected for thermal damage by
the maintenance and operating personnel.

17.3 TESTING COAL BURNERS FOR TIP


VELOCITY

The following is a method to determine the actual tip velocity on coal


burners when no actual flow measurements are possible.

Data Needed
Metric English
W = optimum kiln output, steady state kg/h sh.t/h
H = average specific fuel consumption Id/kg Btu/sh.t
A = fuel heat value (as fired) ldlkg Btu/lb
M = percent moisture in coal (mill inlet) -(decimal)
V = percent volatiles in coal (as fired) -(decimal)
T = primary air temperature C F
p = burner-tip pressure mm HZ0 in. HZ0

Calculations:
1 . Weight of dry fuel jired per minute:

w1 = (Jr!J!!) = kg/min (lb/min)


FUEL SYSTEMS 261

2. Total combustion air required (@ 5% excess air):


a) English units

10.478W = lb air/min

b) Metric units

Ql 10.478W = kg aidmin

3. Primary airflow:

a = WQd = kg/min (lb/r&)

where:
x = %primaryti
Note: In selecting the proper primary air flow, one has to consider the
amount of air needed to evacuate the coal from the mill when the direct-
firing metbad is used. Also important is the fan static-pressure rating.
Guidelines: x = 0.16 - 0.23 for direct-fired kilns
x = 0.07 - 0.14 for semidirect- and indirect-fired kilns

4. Volume of primary air:


a) Determine density (d> of primary air
English units

d = 0.080714 ( TG) (14*7,;0361p) = Ib/ft3

Metric units

d = l.2929(T yi<,) ( 760 ,r3” ) = kg/&

b) Determine volume (VI) of dry primary air


a

a =
Vl = - m?min (ACFM) dry
d
c) Determine the water vapor (~2) from moist coal
(direct-fired systems only)
262 THE ROTARY CEMENT KILN

v2 = wMZ = m?min (ACFM) vapor


where:
Z = density of water vapor (m3ikg, ft3/lb)

Guidelines for Z:
English units Metric units
93 160 180
for ; Ip” 471. 8 1 Fl2 2.06 ;p 77 50 F

Note: The selection of the proper mill-outlet temperature should take into
consideration the volatile content of the coal. As a rule of thumb use:
IfV =>35%, then recommended T = 65 C (150 F)
= 25-35% = 71 C (160 F)
= 12-25% = 82 C (180 F’)
= 5-12% = 90 c (195 F)
= c 5% (petroleum coke) = 102 C (215 F)

d) Determine the actual volume of moist primary air

Q2 = vl + v2 = m3/min (ACFIvI) moist

5. Tip velocity:

Area = 3.1416r
where:
I = radius of burner tip
Q2 =-
VeIocity = - timin (ft/min)
Area
0.016667Q2
= m/s (ft/s)
Area

Using the above method in computing tip velocity, the characteristics


of coal burners have been determined and are shown in Table 17.1 (direct-
fired, wet-process kiln), Table 17.2 (direct-fired, dry-process kiln), and
Table 17.3 (indirect-tired, suspension preheater kiln).
TABLE 17.1
BURNER-TIP VELOCITY ON COAL BURNERS
(Direct-Fired Wet-Process Kiln)
Data Input English Metric

w = optimum kiln output, steady state 32 sh.tfh 29030.4 Wh


If = average specific heat consumption 505000 Btu/sh.t 5873.655 kJ/kg
A = fuel heating value (as fired) 12300 Btu/lb 28622.1 kT/kg
M = percent moisture in coal (mill inlet) 0.11 (decimal) 0.11 (decimal)
v = percent volatiles in coal (as fiid) 0.28 (decimal) 0.28 (decimal)
T = primary air temperature 180 F 82 C
P = burner-tip velocity 5.1 ~in. Hz0 1295 mm Hz0
2
x = percent primary air 0.21 (decimal) 0.21 (decimal) 3
c = diameter of bumcr tip 0.96 ft 10.292 m
effective burner-tip opening 0.7213 ft2 0.0670 m2 Fi

1. Weight of dry fuel fired per minute 2 1 9 . 0 lb/min 99.3 kglmin


2. Total combustion air required 2239.7 lb/min 1016.6 kglmin
3. FJrimary airflow 470.3 lb/min 213.5 kg/m.in
4. Volume of primary air a) density 0.0628 lb/f13 0.8691 kg/m3
b) volume air 7486.5 ACFM dry 245.6 ms/min
c) water vapor 1204.3 ACFM 34.1 ms/min
Total primary air 8690.8 ACFM 279.7 ms/min
5. Tip velocity 12049 ft/min 4174 mlmin
201 ft/s 70 m/s
8
TABLE 17.2 E

BURNER-TIP VELOCITY ON COAL BURNERS


(Direct-Fired Dry-Process Kiln)
Data Input English Metric
W = optimum kiln output, steady state 73 sh.Wh 66225.6 kg/h
11 = average specific heat consumption 4100000 Btu/sh.t 4768.71 kJ&
A = fuel heating value (as fired) 11980 Btu/lb 27877.46 kJfl<g
M = percent moisture in coal (mill inlet) 0.13 (decimal) 0.13 (decimal)
V = pcrccnt volatiles in coal (as fircd) 0.32 (decimal) 0.32 (decimal) ET4
T = primary air temperature 180 F 82 C
= burner-tip velocity %
P 5.6 in. Hz0 1422 mm H20 4
x = percent primary air 0.18 (decimal) 0.18 (decimal) I2
C = diameter of bumcr tip 1.2500 ft 0.3810 m
effective burner-tip opening 1.2272 ft2 0.1140 m2 E
-- 5
1. Weight of dry fuel fired per minute 416.4 lb/mm 188.8 kg/mm
2. Total combustion air required 4148.2 lb/mm 1882.9 kglmin
3. Primary air flow 746.7 lb/mm 338.9 kg/mm
4. Volume of primary air a) density 0.0629 lb/ft’j 0.8569 kg/m3
b) volume air 11870.5 ACFM dry 395.5 m3/min
c) water vapor 2706.5 ACFM 76.6 m?min
Total primary air 14577.0 ACFM 472.1 m%nin
5. Tip velocity 11878 ft/min 4141 mlmin
198 ft/s 69 m/S
TABLE 17.3
BURNER-TIP VELOCITY ON COAL BURNERS
(Indirect-Fired Suspension Preheater Kiln)
Data Input English Metric

w = optimum kiln output, steady state 89 sh.Vh 80740.8 kg/h


II = average specific heat consumption 295oooo Btu/sh.t 3431.145 kJ/kg
A = fuel heating value (as fired) 11980 B m/lb 27877.46 kJk3
M = percent moisture in coat (mill inlet) 0 (decimal) 0 (decimal)
v = percent volatiles in-coal (as fired) 0.26 (decimal) 0.26 (decimal)
T = primary air temperature 180 F 82 C
P = burner-tip velocity 8.5 in. Hz0 2159 mm H,O
x = percent primary air 0.09 (decimal) 0.09 (decimal) 2
c = diameter of burner tip 3
0.67 ft 0.203 m
effective bumcr-tip opening 0.349 1 ft2 0.0324 m2

1. Weight of dry fuel fired per minute 3 6 5 . 3 lb/mm 165.6 kg/mm


2. Total combustion air required 3638.9 lblmin 1651.7 kg/mm
3. Primary air flow 327.5 lb/mm 148.7 kg/mm
4. Volume of primary air a) density 0.0633 lb/f@ 0.7860 kg/m3
b) volume air 5170.2 ACFM dry 189.1 ms/min
c) water vapor 0.0 ACFM 0.0 ms/min
Total primary air 5170.2 ACFM 189.1 m%nin
5. Tip velocity 14811 ft/min 5832 mlmin
247 fvs 97 m/s
b!
WI
18.

Clinker Cooler Control

It is extremely important for the operator to master cooler controls as


well as other control functions of the rotary kiln system, because the
cooler itself performs an integral part of the clinker-burning process.
Cooler conditions influence the burning process in the kiln and conse-
quently, the quality of the clinker. There are many instances where cooler
operating problems can lead a perfectly stable kiln into a severely upset
condition with possible damage to the equipment.
With the introduction of large-sized rotary kilns having production ca-
pacities in excess of 1500 tons per day, traveling grate coolers have
become a problem in some plants. Unexpectedly, widespread complaints
began as cooler components burned up and entire coolers had to be repaired
at alarming rates. These problems were not only experienced in North
America but were also encountered throughout the world wherever large
rotary kilns had been placed into service. Regrettably, a final solution to
the problem has not yet been found, and in some cement plants frequent
cooler failures due to overheating still persist. On the other hand, major
modifications and improvements in the design of cooler equipment in the
past few years have lessened the problems to a large extent and indications
are that coolers can now be made to operate in a satisfactory manner.
Also, not all cooler failures are caused by design of the cooler itself, for
inexperience and laxness on the part of operators has contributed to the
problem. The detailed description of cooler control functions in this
chapter can give the operator the information he needs to help lessen the
frequency of cooler failures.
Fig. 18.1 shows a schematic layout of cooler controls to familiarize the
CLINKER COOLER CONTROL 267

Fig. 18.1 Sectional diagram of a traveling-grate cooler. Instrumentation is


as follows: 1, undergrate pressure, 1st compartment; 2, undergrate pressure, 2nd
compartment; 3, undergrate pressure, 3rd compartment; 4, Fan No. 1 differential
pressure; 5, Fan No. 2 differential pressure; 6, Fan No. 3 differential pressure; 7,
cooler grate temperature; 8, clinker temperature at cooler outlet; 9, secondary air
temperature; 10, grate drive speed; 11, cooler air temperature; 12, hood pressure;
13, cooler outlet gas temperature; 14, excess air chimney; 15, clinker crusher;
16, drive unit; 17, burner.

reader with the various terms used. It does not represent any particular
make of cooler, but shows the components and principles involved in
cooler construction and operation.

18.1 CRITICAL VARIABLES IN


COOLER CONTROL

The kiln operator must operate the cooler in such a manner as to meet
the following objectives as closely as possible:

a) Clinker temperature at the discharge of the cooler should be as


low as possible because high temperatures endanger the transport
268 THE ROTARY CEMENT KILN

equipment and waste valuable heat.


b) Secondary air temperature should be as stable and high as possible
because this is a prerequisite for overall kiln operating stability
and good fuel efficiency.
c) Cooler exit-air temperature should be as low as possible and
volume as small as possible to assure a minimal amount of heat
wasted to the atmosphere.
d) Hood pressure should always be slightly negative.
e) Depth of the clinker bed in the cooler should be such that a free
passage of air through the bed can take place.
f) Cooler control settings should be such that bedgrates, cooler drive
unit, clinker crusher, and cooler walls cannot become overheated.

The kiln operator has basically two control variables for accomplishing
the above-mentioned objectives: the speed of the bedgrates which alters the
clinker residence time and the clinker bed depth in the cooler, or the air
distribution in the cooler can be changed.
In case of an emergency, such as badly overheated cooler conditions, the
kiln operator has two more possibilities to bring the cooler under control:
the amount of clinker falling into the cooler can be decreased by slowing
the kiln speed, or the temperature of the clinker falling into the cooler can
be lowered by adjusting the flame geometry to shift the burning zone
further back in the kiln.
Numerous indicators and recorders on the plant control panel provide the
means by which the operator maintains surveillance over operation of the
cooler and enables the detection of irregularities in operation. These in-
struments should be observed and checked on a regular basis. Those that
are most commonly used are undergrate air pressure, secondary air temper-
ature, grate speed amperage drawn by grate drive motor, and the clinker-
discharge temperature at the cooler outlet. In addition to these five
essential instruments, there are other recorders that assist the operator in
obtaining an overall indication of cooler performance. These additional
instruments record the grate temperature, cooler exit-air temperature or
circulating air temperature, flow rate of air forced into the cooler, and
quench-air temperature measured at a point midway of the cooler length
above the clinker bed. Other frequently used instruments are a nuclear
gauge for measuring clinker bed depth, and a television camera and monitor
showing the cooler interior,
The discussion that follows focuses on traveling grate coolers since
CLINKER COOLER CONTROL 269

these are the most commonly used and the most difficult to control. This
type of a clinker cooler is shown in Fig. 18.1.

18.2 UNDERGRATE PRESSURE AND


AIR-FLOW RATE CONTROL

These two controls, undergrate pressure and air flow, are probably the
most significant parts of cooler control because they constitute the key to
successful achievement of the objective of cooler control. A thorough
knowledge and understanding of these controls is essential to the operator
to enable him to maintain operating stability of the kiln and prevent the
cooler components from overheating.
Before the procedures in undergrate pressure control are explained, a very
important point has to be stressed: The described procedures are only vaild
when the cooler contains clinker that has been properly burned. These pro-
cedures should not be followed when unburned, dusty clinker or raw feed
has entered the cooler.
The cooler system shown in Fig. 18.1 has three undergrate compart-
ments, each compartment receiving cooling air from an individual fan.
The cooler bed-grate drive unit is in the center portion of the cooler. For
control purposes, various instruments record the undergrate pressure in
each compartment, air-flow rates delivered by each fan, and the speed of the
bed grates. Under normal operating conditons (stable operation), there is
an undergrate pressure for each compartment that ensures proper cooling of
the clinker. By holding this undergrate pressure constant, the operator will
theoretically hold the cooler control fairly constant and in balance.
Undergrate pressure is governed mainly by the following factors:

a) Depth of the clinker bed over the grates


b) Average particle size of the clinker in the cooler
c) Temperature of the clinker in the cooler, and
d) Amount of air introduced into the cooler.

Thick beds have higher resistance and thus require more force from the fan
to push the air through than a thin bed. In other words, assuming that the
other factors listed above remain constant, a thicker clinker bed results in
270 THE ROTARY CEMENT KILN

higher undergrate pressure.


Depth of the clinker bed can be-controlled by speed of the cooler grates.
When a thinner bed is required, the grate speed is increased. A deeper bed
can be obtained by slowing the grate speed. Because of the relationship
between undergrate pressure and bed depth, it is possible to maintain a con-
stant undergrate pressure by regulating the grate speed. Some cooler
installations have automatic controls that work on this principle; a set-
point is selected on the controller for a desired undergrate pressure and the
grate speed is automatically adjusted whenever the pressure deviates from
the setpoint.
This approach in cooler control is usually satisfactory when the kiln
operates in a stable fashion. However, during upset kiln conditions, or
when the clinker characteristics change, this approach leaves much to be
desired. During such times, an operator can often experience difficulties in
maintaining proper cooling of the clinker. For example, consider the base
operating condition shown in Table 18.1, with a second compartment bed
depth of 14 in. (36 cm) at an hourly clinker throughput rate of 73 tons. In
this base case, the clinker residence time in the cooler is 20.9 min and the
clinker-discharge temperature 224 F. Now assume that the kiln suddenly
starts to discharge clinker at a rate of 89 tons/h and that the grate speed is
on automatic control, i.e., starts to increase the speed to maintain the same
clinker-bed depth (see Table 18.2). Increasing this grate speed in an
attempt to maintain a constant undergrate pressure, will then result in a
reduction of the clinker residence time in the cooler, as shown in Table
18.2, to 17.2 min. This example shows that by pushing the hot clinker at
a faster rate through the cooler and maintaining the same specific air-flow
rate and undergrate pressure, the clinker-discharge temperature will increase
to a theoretical 317 F. The lesson to be learned from these examples is
that whenever a significantly larger amount of clinker falls into the cooler
and as a consequence the grate drive speed increases significantly, the oper-
ator must compensate for this with a corresponding increase in the specific
air-flow rate (SCFM).
Other influencing factors that cause the undergrate pressure to change
are the temperature of the clinker in the cooler and the amount of air
introduced into the individual undergrate compartments. Assuming other
factors to be constant, a higher temperature of the clinker results in a
higher undergrate pressure and increased air-flow rates in the undergrate
compartments yield higher undergrate pressures.
TABLE 18.1
COOLER OPERATING STUDY (BASE CASE)
Compartment
1 2 3 4 5 Total
Lenga w 1.5 5 12.5 20 24 63
Width (ft) 7 7 9 9 9 E
Depth (ft) 1.50 1.17 0.91 0.91 0.91
T.P.H. output 73 73 73 73 73 f3
56

Weight residence
Clinker of clinkertime
‘ 1544
0.6 4002
1.6 10033
4.1 16052
6.6 19263
7.9 50893
20.9 1
fd
Clinker input “F 2500 2057 1363 640 353 8
Air
Air input
output ““ FF 90
2090 90
1520 90
840 90
420 90 3
Lb air (rcsidcnce time) 350 1980 9800 14ooo 16% 42630 P

Air
Clinker
necdcdoutput
(SCFM) “F 3% 16013
1363 31% 28225
353 27% 110494

MBtu/t heat consumption 4.1


Air needed in kiln (lb/n&) 4077.76 = 54,234.3 SCFM
Excess cooler air (SCFM): 56260.1 = 50.9% of total air in
c:
TABLE 18.2 f:N
COOLER OPERATING STUDY (22% more clinker in cooler but air-flow rate the same)
Compartment

1 2 3 4 5 Total
Length (ft) 1.5 5 12.5 20 24 63
Width (ft) 7 7 9 9 9
Depth (ft) 1.50 1.17 0.91 0.91 0.91
T.P.H. output 89 89 89 89 89

Weight of clinker 1544 .4OO2 10033 16052 19263 50893


Clinker residence time 0.5 1.3 3.4 5.4 6.5 17.2

Clinker output OF 2128 1518 814 494 317


Air needed (SCFM) 7337 16023 31620 28242 27297 110518

MBtu/t heat consumption 4.1


Air needed in kiln (lb/min) 4971.52 = 66121.2 SCFM
Excess cooker air (SCFM): 44397 = 40.2% of total air in
CLINKER COOLER CONTROL 273

18.3 CLINKER RESIDENCE TIME

Clinker residence time in the cooler can be determined on a theoretical


basis when certain factors are known. For the purpose of illustration,
three different cooler operating conditions will now be considered to
determine the effects of changes of these critical variables.
Residence Time at Cosntunt Output. Under these circumstances, the
bed depth is directly proportional to the grate speed Knowing the area of
the cooler, the bed depth under normal operating conditions for a given
kiln output rate, and the density of the clinker, the residence time can then
be calculated by Eq. (18-1).
6Oahy
t= (18-1)
R
where:
t = clinker residence time in the cooler (min)
a = area of the cooler grate surrface (ft2 or m2)
h = beddepth (ft or m)
y = clinker density (lb/fts or kg/ms)
R = biln output (lb/h or kg/h)

Bed Depth at Constant Speed. At constant grate speed, the depth of the
clinker bed in the cooler varies as the kiln output varies. By solving Eq.
(18-1) for h,
W
h =- (18-2)
60~~

In both examples volume in-cubic feet of clinker residing in the cooler


at any time is given by
tR
v= - (18-3)
WY

Residence Time at Conslant Bed Depth. As mentioned previously, it is


not desirable to maintain a constant undergrate pressure under all kiln out-
put rates. This, of course, requires a constant bed depth, which means that
the clinker residence time must be changed as the kiln output changes. Eq.
(18-1) is used.
274 THE ROTARY CEMENT KILN

An operator should have no trouble in finding solutions for the par-


ticular cooler he is operating, as the constants in the preceding equations
can be substituted so that they apply to his cooler and kiln operation, and
can be developed into tables similar to Tables 18.3, 18.4, 18.5, and 18.6.
The reader’s attention is especially directed to Table 18.5 which clearly
shows the extremes mat residence time can span at kiln output rates not
uncommonly obtained during upset operating conditions. It also substan-
tiates the statement that hot clinker-discharge temperatures can become a
reality if insufficient time is allowed to cool the clinker properly.
It should now be clear that proper cooling is not obtained by operating
with a constant undergrate pressure (bed depth) at all times, as it is defi-
nitely wrong to assume that undergrate pressure is a function of bed depth
only. If this were so, cooler components would not bum up at such an
alarming rate as now encountered on many rotary kilns.

18.4 PARTICLE SIZE OF CLINKER

A critical factor in undergrate pressure reacticns is the average particle


size of the clinker in the cooler. The reader tight be familiar with the
Blaine cement surface-testing procedure in which measurement is made of
the time required to pass a specified amount of air through a standard
sample of cement. It takes longer for the air to pass through a fine sample
than a coarse one because the finer material imposes greater resistance
against the air. Exactly the same action takes place when the clinker in
the cooler gets finer because of some upset kiln operating condition. A
finer clinker bed imposes more resistance against air flow, undergrate pres-
sure increases, and the fan has to use more force to push the air through
this kind of a bed than through a normal one. This however is not the
only problem Individual clinker particles are lighter when the particle size
diminishes. Such particles can easily be lifted into the air stream above
the cooler grates because of their lighter weight.
From this it becomes clear that there are two changes in the cooler, one
following the other. First the undergrate pressure increases when the
clinker gets fmer because the smaller particles impede air flow through the
bed. Then, when the air-flow rate is increased to restore the normal flow
of air through the bed, the clinker bed can become fluidized. As scan as
this takes place, resistance of the bed to the flow of air decreases and the
TABLE 18.3
COOLER PARAMETERS:
Clinker residence times at constant kiln output (Grate speed varies)
Kiln Bed Clinker weight Clinker volume Residence
output depth in the cooler in the cooler time
sh.Uh m.ffh in. cm lb kg ft3 m3 min
E
73 66.2 7 18 31657 14361 323.0 9.1 13
73 66.2 8 20 36180 16412 369.2 10.5 15
B
73 66.2 9 23 40702 18464 415.3 11.8 17
73 66.2 Id 25 45225 20516 461.5 13.1 19
73 il
66.2 11 28 49747 22567 507.6 14.4 20
73 66.2 12 30 54269 24619 553.8 15.7 22 b
73 66.2 13 33 58792 26670 599.9 17.0 24
73 8
66.2 14 36 63314 28722 646.1 18.3 26
73 66.2 15 38 67837 30773 692.2 19.6 28 i3
73 66.2 16 41 72359 32825 738.4 20.9 30 8
73 66.2 17 43 76882 34877 784.5 22.2 32
73 66.2 18 46 81404 36928 830.7 23.5 33
73 66.2 19 48 85927 38980 876.8 24.8 35

Length of cooler grates: 63.0 R 19.2 m


Width of cooler grates: 8.8 ft 2.7 m
Grate area: 553.8 ft2 51.4 m2
Clinker density: 98.0 lb/fP 1570.0 kg/m3
Y
VI
TABLE 18.4
COOLER PARAMETERS:
Clinker bed depth variation with changing kiln outputs (Grate speed constant)
Kiln L3ed Clinker weight Clinker volume Residence
011tput depth in the cooler in the cooler time

sh.Vh m.Uh in. cm lb kg ft3 m3 min

62 56.2 10 26 45467 20625 463.9 13.1 22


64 58.0 10 26 46933 21291 478.9 13.6 22 E
66 59.8 11 27 48400 21956 493.9 14.0 22 2
68 61.7 11 28 49867 22621 508.8 14.4 22 2
70 63.5 11 29 51333 23287 523.8 14.8 22 8
22
74 67.1 12 30 54261 24617 553.7 15.7 22
15.3 B
7126 66 85 .. 93 11 22 33 10 55 25 81 03 03 22 35 92 58 23 538.8
568.7 16.1 , 22
78 70.7 13 32 57200 25948 583.1 16.5 22
80 72.5 13 33 58667 26613 598.6 17.0 22
82 74.3 13 34 60133 27279 613.6 17.4 22
84 76.2 14 35 61600 21944 628.6 17.8 22
86 78.0 14 35 63067 28609 643.5 18.2 22

Length of cooler grates: 63.0 ft 19.2 m


Width of cooler grates: 8.8 ft 2.7 m
Grate area: 553.8 ftz 51.4 m2
Clinker density: 98.0 lb/V 1570.0 kg/m3
--.

TABLE 18.5
COOLER PARAMETERS:
Residence time when grate speed is adjusted for changing kiln output rates
(Bed depth is held constant)
Kiln Bed Clinker weight Clinker volume Residence
output depth in the cooler in the cooler time

sh.Vh m.Uh in. cm lb ks ft3 m3 min


&
62 56.2 12 30 54269 24619 553.8 15.7
64 22 65 2
58.0 12 30 54269 24619 553.8 15.7
66 59.8 12 30 54269 24619 553.8 15.7
68 61.7 12 30 54269 24619 553.8 15.7 2 45 i
70 63.5 12 30 54269 24619 553.8 15.7
72 65.3 12 30 54269 24619 553.8 15.7 22 33 b
74 67.1 12 30 54269 24619 553.8 15.7 22 8
76 68.9 12 30 54269 24619 553.8 15.7
78 70.7 12 30 54269 24619 553.8 15.7 21 $
80 72.5 12 30 54269 24619 553.8 15.7 20 ro
82 74.3 12 30 54269 24619 553.8 15.7 20
84 76.2 12 30 54269 24619 553.8 15.7 19
86 78.0 12 30 54269 24619 553.8 15.7 19

Length of cooler grates: 63.0 ft 19.2 m


Width of cooler grates: 8.8 A 2.7 m
Grate area: 553.8 ftz 51.4 rnz
Clinker density: 98.0 Ib/fP 1570.0 kg/m3 2
TABLE 18.6
2
COOLER OPERATING STUDY
(same as base case but higher bed depth, kiln output and total air into cooler the same)
Compartment

1 2 3 4 5 Total
Length (fl) 1.5 5 12.5 20 24 63
Width (ft) 7 7 9 9 9
&Pth (ft) 1.83 1.33 1.03 1.03 1.03
T.P.H. output 73 73 73 73 73
2
Weight of clinker 1883 4562 11356 18169 21803 57773 2
Clinker residence time 0.8 1.9 4.7 7.5 9.0 23.7
B
Clinker
Air inputinput “OFF 2500
90 2137
90 1537
90 940
90 741
90 $
Air output “F 2090 1500 790 350 222
Lb air (residence time) 350 1980 9800 14000 16500 ’ 42630

Clinker
Air needed
output
(SCFM) “F iE 14047
1537 27930
940 24937
741 2472 97052

MBtu/t heat consumption 4.1


Air needed in kiln (lb/mm) 4077.76 = 54234.3 SCFM
Excess cooler air (SCFM): 42817.7 = 44.1% of total air in
CLINKER COOLER CONTROL 279

undergrate pressure will suddenly drop. A fluidized clinker bed is a highly


undesirable and dangerous condition because a bed in such a state usually
does not move along properly in the cooler. On a horizontal grate cooler
the clinker bed when fluidized tends to remain stationary with more and
more clinker buiIding up on it. Then when sufficient weight has been
acquired by the bed and it starts to move again, there is so much clinker
present in the cooler that it cannot be properly cooled, and it could again
choke off air flow through the bed, starting the cycle over again.
It is a fundamental rule of traveling grate cooler operation never to
permit raw feed or extremely fine clinker to enter the cooler. The most
modem traveling grate cooler in use today is not designed to handle such
fine materials and can become overheated and damaged if called upon to do
so. The best procedure in such an instance is to lower the kiln speed in
order that the load in the cooler is lessened, thus giving enough time for
the fine clinker to cool properly.
Another question which has to be addressed is: “How does the clinker-
discharge temperature react when the kiln output remains the same but the
clinker bed depth in the cooler is increased?” Such an increase in bed depth
is usually the result of:

a) deliberate decrease in grate speed due to the selection of a higher


undergrate-pressure setpoint or
b) a design change where the effective grate area of the cooler is
reduced.

Naturally, this increase in bed depth (undergrate pressure) immediately im-


poses a higher restriction on the cooler air fans and demands sufficient fan
capacity to overcome this added restriction. Consider the base case (Table
18.1). Here, 42,630 lb of air are given to the cooler in 20.9 min to cool
50,893 lb of clinker. Assume that this cooler would be operated with a
higher bed depth but-with the same kiln output and 42,630 lb air input.
This operating condition is shown in Table 18.6 and indicates that the
clinker residence time has increased from 20.9 to 23.7 min. due to the
increase in residence time, the 42,630 lb of air now represents a specific air-
flow rate of only 97,052 SCFhI. The most important aspect of this
change is that the resultant increase of the clinker-discharge temperature
from 224 F to 641 F is completely unacceptable and dangerous. Here too,
the solution is to increase the specific air-flow rate (SCFh4) to compensate
for the increase in bed depth.
280 THE ROTARY CEMENT KILN

18.5 OPERATION OF COOLER FANS

Now consider control of the rate of air flow in the cooler. The prime
requirement in cooler control is to make sure that the air flow through the
clinker bed is never restricted completely, because such a restriction leads
directly to insufficient cooling of the clinker and possible damage to the
cooler components. It has been pointed out that different clinker beds exert
different resistances against the cooler fans. .To be able to understand this
important fact more clearly, the operator must have some knowledge of
how a fan works under actual operating conditions.
Fig. 18.2 is a simplified illustration of an undergrate compartment with
its corresponding air fan. Air-volume control for cooler fans is most com-
monly caked out by means of either fan outlet dampers, fan inlet vanes,
or sometimes both. The fan speed is constant for any one cooler fan SO it
is necessary for the operator to change the position of the damper or vane
to reduce or increase the volume of air moved by the fan assuming no
change in static pressure on the fan.

UNDERGRATE
PRESSURE
RECORDER

UNDERGRATE

COMPAlTMENI

DIFFERENTIAL
PRRSSURE
RECORDER

Fig. 183 Schematic diagram of one undergrate compartment of a traveling-


grate cooler.
cLINKJzcooLERcoNTFtoL 281

Fan manufacturers usually supply performance curves of individual fans


for the customer. Fig. 18.3 is such a curve showing the volume of air
plotted against horsepower and static pressure. In order that this discussion
does not become too technical and enter fields over which an operator has
no control, the horsepower curve will be disregarded and attention directed
to the static pressure curve.

FAN HP

2 4 6 0 10 12
FAN STATIC PRESSURE IN.OF WATER

Fig. 18.3 Typical fan performance curve.

Fan static pressure is the total pressure developed by the fan, less the
velocity pressure in the fan discharge duct. For all practical purposes, fan
static pressure in a cooler installation is equal to the undergrate pressure,
within close tolerances. Air flow is a function of static pressure and power
T-HE ROTARY CEMENT KILN

applied to the fan. To make this discussion more meaningful, substitute


undergrate pressures for equal units of fan static pressures, and restrict the
range of pressures to values as they are encountered normally on rotary-
kiln coolers, thus obtaining new fan performance curves as shown in Fig.
18.4. The curve on the right represents the same curve as shown in Fig.
18.3, that is, the dampers of the fan are wide open, and the air flow is the
maximum obtainable for this particular fan. This curve shows clearly how
the air volume decreases as the undergrate pressure increases. For example,
when the undergrate pressure is 9 in. of water with the damper wide open,
the volume of air moved by the fan is 24,000 CFM. If the undergrate
pressure increases, perhaps as a result of increased kiln output, the volume
of air progressively decreases until at 12 in. undergrate pressure the volume
is only 20,ooO CFM. With the fan dampers wide open, an increase in
undergrate pressure causes less air to pass through the clinker bed, creating
the dangerous possibility of an overheated cooler and at the same time
insufficient cooling of the clinker.

UNDERGRATE PRESSURE IN.DF WATER

Fig. 18.4 hfomxincc curve for a fan under different inlet vane openings.
Operation is apt to be critical and unstable at pressures above 12 in. for this
paticular fan.
CLINKER COOLER CONTROL

During normal and stable operating conditions, the fan dampers or


vanes are never fully open. In Fig. 18.4 additional curves have been
plotted showing inlet vane openings of 90, 80,70, and 60%. Usually, a
cooler fan is operated at approximately 60% opening, thus giving the
operator the necessary freedom to increase air flow if a kiln upset should
make it necessary. For example, assume that the cooler is operating in a
stable fashion with adequate cooling of the clinker at an undergrate pressure
of 7 in., fan inlet vane at 60% opening, and air volume of 21,000 CFh4.
For some reason, the undergrate pressure increases to 11 in. In order to
obtain the same volume of air through the clinker bed, the fan inlet vane
opening must be increased from 60% to 90%. If the vane were not
opened, the air volume would drop to 19,000 CFM.
These examples of fan performance under actual operating conditions
point out two significant facts: First, as with any butterfly-type valve or
damper, maximum flow is reached when the damper is about 88% open.
Any enlargement in the opening thereafter gives only a very small increase
in air flow. Second, air-volume output is directly related to undergrate pres-
sure. Therefore, the fan must have sufficient capacity to provide the
necessary amount of air at the maximum anticipated undergrate pressure.
A word of caution must be introduced here. Each fan installation will
have its own characteristics. The fan curves discussed in this chapter apply
to one certain installation and can be considered as typical, to show the
reader the relationships between pressure, output, and power requirements.
Reactions and capacities of any individual installation must be computed
for that particular combination of equipment. For this reason, it is advis-
able for an operator to examine the curves of the particular cooler fans that
are under his control, and become familiar with their characteristics and
capacities, so he will know the operating limits for undergrate pressure.
Whenever the undergrate pressure exceeds these limits, the operator will
know then that less efficient cooling is taking place and that an immediate
change in kiln speed must be made to avoid damage to the cooler.

18.6 CLINKER AND AIR DISTRIBUTION


IN THE COOLER

Finally, it is necessary to consider distribution of clinker and air in the


cooler. For proper cooling of the clinker it is essential that the clinker is
evenly spread over the width of the cooler so that the bed offers uniform
THE ROTARY CEMENT KILN

resistance to the passage of air throughout its width When the clinker
passes to one side of the cooler, leaving a thinner bed on the opposite side,
the air will naturally seek a passage through the bed where it offers the
least resistance. Consequently, the air passes through the bed where it is
least needed and little air passes where it is needed’most. Formation of
stalagmites (commonly referred to as “snowmen” or “candles”) at the cool-
er inlet is the prime cause of this condition. Various devices are used to
combat stalagmite formations. Some coolers have watercooled steel
jackets, or watercooled clinker spreaders, and others have a special row of
quench grates with their own air supply, to spread the clinker rapidly over
the width of the cooler at the inlet.
For proper distribution of air in the cooler no air should freely pass
from one undergrate compartment to another through large leaks or other
openings. If this is allowed to take place, the air introduced, for example
into the first compartment, could pass over into the second compartment
when the clinker bed is thicker at the cooler inlet.
Proper clinker distribution is an acute problem in the upper region of
many grate coolers. This problem can usually be overcome by:

a) narrowing the cooler width by means of installing refractory


ledges and or “dead” grates along the cooler walls and
b) concurrently increasing the static-pressure capability of the air
fans in these narrowed compartments.

In Chapter 12 it was briefly mentioned how important it is to apply the


maximum air in these compartments where it can do the most thermal
work This means, the majority of the cooling work should be accom-
plished in the first and second compartments because it is here that the
most efficient heat transfer will take place due to the large temperature
difference between the air and the clinker. Once again, the base case in
Table 18.1 is considered and it is assumed that 2000 lb more air is added
into the fourth cooler compartment, i.e., at the place in the cooler where
heat transfer is relatively inefficient. Table 18.7 shows that this 4.7%
increase in total air input results in a clinker-discharge-temperature drop of
only 13 F (from 224 F to 211 F). It is assumed that all compartment air
flows are readjusted to take advantage of the aforementioned thermo-
dynamic principle. In Table 18.8 air has been added to the second and third
compartments while it has been reduced to the fourth and fifth compart-
ments. The end result is that the total air input into the cooler has been
TABLE 18.7
COOLER OPERATING STUDY
(same as base case but 2000 lb more air into fourth compartment)

1 2 3 4 5 Total

Length (ft) 1.5 5 12.5 20 24 63


Width (ft) 7 7 9 9 9 I?
DePtJl (ft) 1.50 1.17 0.91 0.91 0.91
T.P.H. output 73 73 73 73 73 8

Weight of clinker 1544 4013 10033 16052 19263 50905 ii


Clinker residence time 0.6 1.6 4.1 6.6 7.9 20.9
b
Clinker input OF 2500 2057 1365 642 355 8
Air input OF 90 90 90 90 90
Air output OF 2090 1520 840 420 9
Lb air (residence time) 350 1980 9800 14000 18% 44630 8

Clinker output “F 1365 355 211


Air needed (SCFM) 15%8 31:: 28225 30614

MBtu/t heat consumption 4.1


Air needed in kiln (IWmin) 4077.76 = 54234.3 SCFM
Excess cooler air (SUM): 59524.2 = 52.3% of total air in
TABLE 18.8
COOLER OPERATING STUDY
(same as base case but adding more air in second and third,
and reducing air in fourth compartments1
Compartment

1 2 3 4 5 Total
Length (ft) 1.5 5 12.5 20 24 63
Width (ft) 7 7 9 9 9
&Pth (ft) 1.50 1.17 0.91 0.91 0.91 E
T.P.H. output 73 73 73 73 73 j!
Weight of clinker 1544 4013 10033 16052 19263 50905 2

Clinker residence time 0.6 1.6 4.1 6.6 7.9 20.9 B


Clinker input OF 2500 2057 1221 498 300 2
Air input “F 90 90 90 90 90
Air
Lb airoutput
(residence time) “F 2090
350 2425
1500 10500
790 12300
350 112Z 36575

Clinker output “F 2057 1221 498


Air needed (SCFM) 7339 19556 33871 103767

MBWt heat consumption 4.1


Air needed in kiln (lb/mm) 4077.76 = 54234.3 SCFM
Excess cooler air (SCFM): 49532.3 = 47.7% of total air in
CLINKER COOLER CONTROL

reduced by 15% but the clinker-discharge temperature remains the same at


224 F. Once again, it is important for an operator to remember this
principle of cooler-air application as this will lead to efficient cooler
operation and possibly to less failures and damages to cooler components.
Undergrate pressures are usually set so that the highest pressure is found
in the first compartment and the lowest in the last compartment. Clinker
temperatures above the grates, as pointed out earlier, influence the under-
grate pressure. This step-by-step lowering of the undergrate pressures from
one compartment to another is a natural result of the succeeding lower
temperaures of the clinker as it travels down the cooler.

18.7 SECONDARY AIR-TEMPERATURE


CONTROL

Secondary air temperature has a direct influence on the geometry of the


flame and the point of ignition of the fuel, consequently irregular
secondary air temperatures can cause irregular flame characteristics which
in turn can cause a shifting of the burning zone. Stable kiln conditions are
practically impossible as long as the secondary air temperature is not held
constant, within a tolerance of + 100 F (40 C). The following discussion
applies to controls to be exercised when the kiln is operating at normal
speed and is producing well-burned clinker.
The kiln should be operated with the secondary air temperature as high
as possible because the maximum amount of heat is then recovered from
the clinker, thus improving fuel efficiency, as less fuel is required to raise
the temperature of the air entering the kiln. This condition maximizes
kiln capacity. A second advantage of high secondary air temperature is the
favorable influence on the objective of burning the clinker close to the
nose (front) of the kiln. However, there are practical limits for secondary
air temperatures. A temperature that is too high can result in overheating
in the kiln nose and the burner hood as well as in the cooler. It is
advisable to operate the kiln with the secondary air temperature slightly
below the maximum allowable in order to protect the bed grates and cooler
refractory from damage.
The two factors having the greatest influence on secondary air tempera-
ture are speed of the bed grates in relation to the volume and temperature of
air introduced into the cooler to cool the clinker, and temperature and size
THE ROTARY CEMENT KILN

of clinker discharging from the kiln into the cooler. Under stable kiln con-
ditions, the secondary air temperature is controlled by the speed of the
cooler grates which merely means that the depth of the clinker bed in the
cooler is the controlling factor. Thus, an increase in grate speed (lessened
bed depth), other conditions remaining unchanged, results in a lower secon-
dary air temperature, and a slower grate speed (thicker bed depth) causes an
increase in temperature. It is important to remember, however, that secon-
dary air temperature control is not merely a matter of speeding up or slow-
ing down movement of the bed grate. As mentioned earlier, there are
several factors to be considered before deciding which adjustment will give
the desired results.
Because the secondary air temperature is controlled mainly by the bed-
grate speed and the volume and temperature of the air in the cooler, as well
as the temperature and size of the clinker, this control goes hand in hand
with undergrate-pressure control, as any change in these two variables will
also change the undergrate pressure. This then will considerably limit the
extent to which secondary air temperature can be controlled.
Now consider for example, an upset kiln condition in which the greatest
part of the material in the cooler is in the form of very small-sized nod-
ules, or even worse, in the form of dust. In a situation like this the opera-
tor will first reduce the kiln speed, which reduces the amount of clinker and
consequently lowers the secondary air temperature (see Table 18.9). If an
attempt is made to hold the secondary air temperature within a loo-degree
tolerance, it would be necessary to slow down the bed-grate speed to such
an extent that it would choke off the free passage of air through the bed.
The clinker would thus not be properly cooled and would probably be red
hot on leaving the cooler, causing considerable damage to the clinker trans-
port equipment. The proper adjustment in this case would be to slow
down the bed grates only to such an extent that the normal bed depth can
be maintained. It is also extremely important that the air flow into the
cooler be increased in order that the clinker can be properly cooled. The
operator should never attempt to hold the secondary air temperature at its
normal level when the kiln has been slowed down because of an upset.
The other factor to be considered in secondary air-temperature control is
the temperature of the clinker as it discharges from the kiln. By changing
the character of the flame, the burning zone can be shifted closer to the
kin front thus raising the clinker-discharge temperature and consequently
the secondary air temperature, or the burning zone can be shifted further
back in the kiln, reducing the secondary air temperature. These actions are
summarized in Table 18.9.
CLINKER COOLER CONTROL

TABLE 18.9
SECONDARY AIR-TEMPERATURE CONTROLS

To reduce secondary air temperature: To raise secondary air temperature:

Increase- cooler grate speed. Decrease cooler grate speed.


Decrease clinker bed depth in cooler. Increase clinker bed depth in cooler.
Decrease size of clinker particles. Increase size of clinker particles.
Decrease amount of clinker. Increase amount of clinker.
Bum further back in the kiln. Bum closer to the nose (front) of the
kiln.

Earlier in this chapter it was pointed out that more air must always be
given to the cooler than what is needed for combustion in the kiln. Hence,
a certain amount of excess air has to be vented to the atmosphere or used
for the drying and grinding of raw materials and/or coal. On precalciner
kilns, this excess air is diverted to the flash calciner by means of the ter-
tiary air duct. By inference there must therefore be an imaginary dividing
line in the cooler wherein the air in the hotter region goes to the kiln and
the one in the colder part moves toward the cooler stack. Simple calcu-
lations on the operating conditions explained in Tables 18.1 and 18.8
would show that more efficient apphcation of the air in the first two
compartments would result in a shift of this imaginary line toward the
upper end of the cooler, higher secondary air, and lower cooler-stack air
temperatures.
Design limitations, either in cooler size or fan capacities, often lead to
excessively high clinker-discharge temperatures regardless of how well the
air distribution is controlled or the cooler mechanically maintained. Such
kilns are usually capable of producing more clinker but the cooler acts as
the bottleneck toward these higher output rates. Short of installing an
after-cooler (such as the well-known G-cooler) or major modifications in
cooler design, there is not much one can do to overcome these limitations.
Some plants use water-spray cooling of the clinker within the last cooler
compartment (usually leading to operating problems with the dust collec-
tor) or make use of reciprocating clinker “skips” combined with water
sprays after the clinker discharges from the cooler. These solutions must
be viewed as only temporary as most of these create additional operating
nrnhlmnr
THE ROTARY CEMENT KILN

18.8 HOOD-DRAFT (PRESSURE) CONTROL

A definitive volume of air is drawn into the kiln by the I.D. fan and a
given amount of air is forced into the cooler by the cooler fans. Years
ago, on wet-process kilns with their high specific-heat consumption and
relatively low kiln output, these two flows were nearly equal. Hence, only
small amounts of excess air had to be vented to the atmosphere by means
of the cooler stack. Some of these kilns were equipped with so-called
closed-circuit cooler-air systems (recycling excess air to the upper cooler
compartment fans) and successfully eliminated all excess air. However, as
kilns became more efficient in heat consumption and clinker output rates,
excess cooler-air volumes started to increase and in many cases this led to
difficulties in hood-draft control. All~too often the damper in the cooler
stack which regulates the hood draft is found to be in the fully open
position thus eliminating any effective control of the hood draft.
Hood-draft control is simply a regulation of the amount of excess air
which escapes through the cooler chimney, so that when the hood pressure
is too high, the damper must be opened, and when the hood pressure is too
low, the damper must be closed. This does not mean that the damper has
to be fully closed or fully opened, but instead small adjustments in the
damper position can be made to give the desired results.
As mentioned, hood draft is governed mainly by:

a) the I.D. fan speed and


b) the volume of air given to the cooler fans.

Assuming that other factors remain constant, then an increase in I.D. fan
speed results in a lower hood pressure, and a reduction in I.D. fan speed
results in a higher hood pressure. Similarly, increasing the amount of air
forced into the cooler results in higher hood pressure.
The hood pressure can be either negative or positive. A negative pres-
sure indicates that the hood is under a vacuum, and a positive hood pres-
sure means that the hood is pressurized. The basic rule in hood-pressure
control is never to operate a kiln with positive hood pressure because this
results in troublesome kiln operating conditions. Fine clinker particles are
blown through the nose-ring seal thus causing the seal to wear out pre-
maturely, and dust emission in the hood area can make viewing of the
burning zone by the operator unpleasant and unsafe. On rotary kilns
CLINKJZR COOLER CONTROL

equipped with optical pyrometers and television cameras for burning-zone


control, positive hood pressures could lead to damage to this equipment
from the flying hot particles. Formation of rings and of stalagmites
(“snowmen”) in the cooler inlet can be attributed to positive hood pres-
sures on some kilns. The importance of operating a kiln with negative
hood pressure is obvious.
There is one situation in which the operator has to take exception to the
above rule: Whenever prevailing high cooler temperatures could cause
damage to the cooler components, it is necessary to introduce a sufficient
amount of air into the cooler to lower the temperature and overcome the
dangerous situation. The first corrective action in such an instance would
be to slow down the kiln speed to lessen the load in the cooler. Then if
dangerous overheated conditions still prevail and the damper at the chimney
is already wide open, one has no other choice but to introduce an added vol-
ume of air into the cooler. However, the operator should return the hood
pressure to negative again at the earliest possible time as soon as the
situation has been brought under control.
The hood pressure is usually automatically controlled. The controller
receives an input signal from the hood-pressure measuring instrument and
sends an output signal to the damper at the cooler chimney. The operator
adjusts the setpoint on the controller to the desired hood pressure (generally
between -0.07 and -0.03 in. of water). Any time the hood pressure
deviates from this setpoint, the chimney damper will then be automatically
adjusted by the controller.
19.

Kiln Exit-Gas Temperature Control

The technique of burning a kiln from the rear was briefly mentioned in
Chapter 14. In this technique, equal or more consideration is given to the
so-called kiln back end (temperature and oxygen content) than to the
burning zone because proper and stable back-end conditions lead to stable
burning-zone conditions. Full control must also consider the pressure or
draft conditions at the rear of the kiln, as the draft pressure is an indication
of the presence of several irregularities that affect operation and control of
the kiln.

19.1 BACK-END TEMPERATURE

Whenever the term “back-end temperature” is used it refers to the exit-


gas temperature in wet and dry kilns without chains and to the intermediate
gas temperature in kilns equipped with chains. For preheater and pre-
calciner kilns, back-end temperature herein refers to the rotary-kiln exit-gas
temperature, i.e, the gas conditions at the inlet to the lowest stage of the
preheater cyclones.
A tight back-end temperature control is the key to a successful, stable
kiln operation. There is sound reasoning behind the procedure of control-
ling the back-end temperature within a narrow range: kiln feed while travel-
ing through the kiln undergoes physical and chemical changes all of which
are important and cannot be overlooked if the kiln is to operate in as stable
a condition as possible. The most crucial change affecting the feed takes
KILN EXIT-GAS TEMPERATURE CONTROL

place in the calcining zone. Froper burning of the’feed in the burning zone
cannot take place unless the feed is completley calcined before it enters the
burning zone. In other words, clinkerizing will not proceed until all the
carbon dioxide (CO2) has been driven off the feed. It should now be clear
that a certain heat gradient has to exist throughout the kiln in or&r that
drying and calcination proceeds in the desired manner. There must .be
means of controlling calcination and drying of the feed before it enters the
burning zone. Back-end temperature control is the means by which the
operator keeps control over feed preparation behind the burning zone.
The fact that the operator controls the back-end gas temperature instead
of the feed temperature (whenever both are measured) raises the question of
whether the feed temperature in the kiln is more important than the gas
temperaure. This of course is true, but the important fact is that the gas
temperature is easier to control than the temperature of the solids.
On dry- and wet-process kilns, the solids temperature does not react as
fast as the gas temperature whenever a change in the control settings has
been made. By controlling the gas temperature, the operator can observe a
change minutes after an adjustment has been made, but considerable time
could pass before a reaction in the temperature of the solids would be
observed. Nevertheless, it would be wrong to assume that the solids
temperature at the back end of the kiln is not important. This temperature
does reveal considerable information in regard to the heat transfer taking
place in the chains.
Ideal Condirions. For any kiln speed and feed rate on a kiln, there is an
ideal back-end temperature that will ensure proper preparation of the feed.
If too much heat exists toward the rear of the kiln, the feed will un-
necessarily undergo early completion of calcination and clinker formation
will start further back into the kiln. Aside from shifting the burning zone
further back, too much heat to the rear of the kiln represents poor oper-
ating efficiency and results in a waste of fuel. In the opposite direction,
not enough heat to- the rear of the kiln could cause the undesirable
condition in which feed is not completely calcined when it enters the
burning zone, rendering the bumability extremely difficult and probably
leading to an upset in the operation.
The back-end temperature is governed by several factors. Of these, the
I.D. fan speed and the fuel rate are the usual causes for changes in back-end
temperature. Whenever one variable is changed and all others remain
constant then higher back-end temperature will result if there is an increase
in I.D. fan speed, an increase in fuel rate, or a decrease in feed rate.
THE ROTARY CEMENT KILN

Conversely, lower back-end temperature results from a decrease in I.D. fan


speed, a decrease in fuel rate, or an increase in feed rate.
Changes in feed rate usually do not happen in a mechanically-sound
rotary kiln if the kiln speed remains constant. Nevertheless, from time to
time conditions will arise in which the feed loading of the kiln changes due
to mechanical failure of the feeding system, feed holdback in the heat
exchangers (chains, preheater cyclones), or most commonly, alteration of
the kiln speed. This in turn will have a direct influence on the back-end
temperature.
To summarize, changes in the following variables can lead to changes
in the back-end temperature:

a ) I.D. fan speed (kiln-exit draft)


b ) fuel rate in the burning zone
c) feed rate
d ) kilnspeed

Clearly, this again points out an important aspect of kiln control, namely
that changes in fuel rate, kiln speed, and/or I.D. fan speed cause changes at
both ends of the kiln (burning zone and back end). Hence, if one changes
any of the above-mentioned kiln-control variables, one has to consider the
effect this change might have throughout the entire kiln. For example, if
the back-end temperature is low one can not take a simple approach and
just increase the I.D. fan speed without considering the effect this
adjustment has on the burning-zone temperature and oxygen content in the
exit gases.

19.2 BACK-END DRAFT CONTROL

Back-end draft, or any kind of draft measurement on the kiln system, is


measured by means of a sensing tube inserted into the gas stream (see
Chapter 12). For back-end draft, one measures predominantly the local
static pressure. Since the kiln operates under induced draft by the large fan
at the kiln back end (hence the name I.D. fan) this local pressure is nega-
tive, i.e., the suction is indicated by the instrument. This is important to
remember because it explains the reasons why this instrument reading
either increases or decreases under certain kiln conditions. If there is a
KILN EXIT-GAS TEMPERATURE CONTROL

restriction downstream (further down the kiln) from this measuring point,
this “pressure” increases because there is more suction. However, if the
restriction is between the I.D. fan and the point of measurement (i.e.,
upstream) this suction decreases.

a) Long dp and wet-process kilns.

In the burning technique previously described, the operator uses the


back-end draft pressure recorder as an indicator of ring formation, slabbing
of rings, feed holdbacks in the chain section, irregularities in the I.D. fan
performance, or for the detection of possible air leakage in the system
between the kiln rear and the I.D. fan. The back-end draft should not be
used for regulating the temperature profile in the kiln because too many
factors can cause a change in this draft. In or&r to achieve satisfactory
kiln operation, the main objectives should be a constant back-end tempera-
ture, an oxygen range of 0.7-1.5%, and a secondary air temperature as
nearly constant as possible.
Some plants have attempted to control ID. fan speed in closed-loop
automatic-control mode by using the back-end draft as the sole input signal
to the controller. In other words, the I.D. fan would only change if there
would be a change in back-end draft. Such efforts fail in most instances
for the reason that the I.D. fan is not able to provide the kiln with the
constant air flow rate that was wrongfully assumed. This also ignores the
temperature profile of the kiln, the rings, or other areas where buildup
might take place. The same applies to attempts that have been made to
control the I.D. fan speed based on an input signal from the percent
oxygen in the exit gases. Constant oxygen levels, in theory, assures a
constant .fuel-to-air ratio. This is a great concept for the internal com-
bustion engines but has little value for rotary-kiln control. Both these
concepts might have some limited application possibilities during stable
kiln operation but are definitely counterproductive during upset conditions.
The truism can again be repeated: “Kiln burning is not so much a task of
maintaining stable operation but rather a specialized skill of knowing how
to prevent upsets and how to stabilize a kiln after an upset.” Kiln
operators are being measured by these skills with the same criteria
applying to the control program that is stored in the memory in the
computer.
Changes in the I.D. fan speed are the most likely causes of changes in
the back-end draft. This is only natural because with every alteration in
THE ROTARY CEMEW KILN

the ID. fan speed the velocity of the gases passing through the kiln
changes. Hence, whenever the operator detects a change in the back-end
draft his first question should be: “Was the I.D. fan speed altered just prior
to the time this change in the draft was observed?” If the answer is
negative, then he can most likely find the reason in one of the following:

1. an inspection door between the kiln and the I.D. fan has been
opened or closed;
2. the fuel chamber door has been opened or closed (as for example
when the dust from this chamber has to be emptied out into a
truck); or
3. the cold air breed-in damper position in the kiln rear has been
changed (wherever such a device is in existence due to I.D. fan
inlet temperature restrictions).

If any one of the above-mentioned doors or dampers is opened, the back-


end draft will decrease because atmospheric air is being pulled into the
system, reducing the amount of gases pulled through the kiln itself. On
the other hand, closing of any of the above doors will cause the draft to
increase. Depending on the size of the door or the damper, changes in the
draft can be either very small or so large that they could lead the kiln into
‘an upset if not counteracted at once.
Losing kiln draft in any of the ways described will lower the oxygen
content in the flue gases and will also lower the back-end temperature.
Under such conditions, it is advisable to adjust the I.D. fan speed to return
the kiln draft to its normal level, which should also ensure a return to the
normal oxygen and back-end temperature. This illustrates an important
usage of the back-end draft recorder, as it indicates immediately any change
in the gas velocity inside the kiln, which under stable kiln conditions
should be fairly constant.
Any workman who has to open a door at the back end of the kiln must
always report his intentions to the operator first, obtaining the operator’s
permission before making any changes. It certainly does not help the
operator any if someone changes the kiln draft without his knowledge,
especially if the kiln is already on the brink of an upset.
Now assume that there has been a drastic drop in the kiln draft but there
has not been a change in fan speed, nor have any doors or dampers been
opened. In this case the origin of the change must be sought in the kiln
itself. Whenever the I.D. fan speed is constant and an obstruction, such as
KILN EXIT-GAS TEMPERATURE CONTROL

a ring in the kiln, is willfully or accidently removed, there will be an


immediate drop in the back-end draft. On the other hand, if a ring is in the
process of forming, the back-end draft will increase progressively as long
as the ring keeps on building up. This is another important reason that
justifies a close watch of the back-end draft Ring formation or droppage
can be easily and promptly detected by evaluation of the draft recording
chart as shown in Fig. 19.1. This is especially helpful in case a ring has
broken from behind the burning zone or from the chain section, giving the
operator time to adjust the kiln settings for the expected heavier load which
will arrive in the burning zone.

DRAFT

Fig. 19.1 A potion of a backend draft nxorder chart, showing: 1, normal


draft; 2, a gradual and steady increase in draft pressure indicates that a ring is
forming; 3, the ring is stating to disintegrate and pressure returns to normal
when the ring breaks loose; 4, a sudden drop in pressure indicates that a door or
louvre has been opened; 5, the door or louvre has been closed and the draft
pressure returns to normal at 6.
THE ROTARY CEMENT KILN

b) Preheater and Precalciner Kilns.

In contrast to conventional long kilns, operators of preheater and


precalciner kilns have the added responsibility of closely monitoring the
draft conditions in each preheater vessel and, in the case of the precalciner
kiln, the draft in the tertiary air duct also. The so-called pressure drop
across a preheater stage is important because it gives the operator an
indication of buildup problems within the vessel or discharge chutes.
Whereas a ring might build up relatively slowly within a wet-process kiln
over a period of a few days, an alkali-sulfate or calcium-sulfate buildup in
the feed chute or lower preheater stage can progress very rapidly and,
consequently, can produce large changes in draft in a short period of time.
20.

Feed-Rate Control

Most modern rotary kilns have feed-metering devices which, being


synchronized with the kiln speed, maintain a uniform depth of feed bed
throughout the entire range of possible kiln speeds. The cross-sectional
loading of the kiln remains theoretically the same regardless of speed of
rotation, as long as the feed-to-kiln speed ratio is not changed
An element of feed rate that has to be considered is the amount of dust
which is returned into the kiln, a condition that is overlooked in many
cement plants. In extreme cases, dust is returned to the kiln at the same
rate as it is collected in the dust-collecting units. This means that for any
given time period there could be very little, and at other times there could
be a large amount of dust returned to the kiln. Although the feed rate is
synchronized with the kiln speed, the irregular dust return will cause an
uneven loading in the kiln. This problem can be partly overcome by
installing a surge bin which to some extent equalizes large variations in
the dust-collecting rates. This procedure still does not feed the kiln in the
same fashion as the kiln feed because the dust return is not synchronized
with the kiln feed. This being the case, the operator should, if he has the
means available, make an attempt to synchronize the dust rate manually
with the kiln speed whenever the kiln speed is changed. If the kiln speed
is decreased, less dust should be returned to the kiln. Conversely, more
dust should be returned to the kiln whenever the kiln speed is increased.
Now assume that this problem does not exist and raw feed as well as
dust is fed to the kiln at a synchronized rate. One could therefore conclude
that the kiln would be evenly loaded from the inlet to the outlet. If this
held true at all times, operation of a rotary kiln would be greatly sim-
THE ROTARY CEMENT KILN

plified. This, however, is not the case. Feed can flush at great speed into
the burning zone because of erratic calcining conditions, or feed can be
retained in the chain section, or in the rear of the burning zone as a result
of ring formation. When the ring breaks loose (slabbing), there will be a
sudden surge of feed into the burning zone to be clinkerized. Any ring in
any part of the kiln will act as a dam obstructing the uniform advancement
of the feed through the kiln.
It is important that the kiln operator keep a close watch over all instru-
ments registering feed loading or feed advancement in the kiln, so that any
abnormalities can be noted as they develop. As with all other control
functions, early detection is of prime importance because the earlier any
irregularity is detected and counteracted, the smaller the chance of leading
the kiln into a serious upset, with the resulting damage to the equipment.
Chains, coating, and even refractories have been lost (melted) because of
overheated conditions resulting from feed shortages and negligence on the
part of the operator. Uniform feeding of the kiln is the first step toward
stable kiln conditions.
When a kiln operates for a prolonged period of time in a stable fashion,
there inevitably arises the question of whether the production rate of the
kiln can or should be increased. Unfortunately, the answer to this question
is often dictated by prevailing clinker inventories and existing commit-
ments to cement sales. These conditions often impose higher output
demands on the kiln at the expense of operating stability and ease of con-
trol. However, despite their common objections kiln operators must learn
to cope with these factors.
It should not be assumed that a kiln may be increasingly “force-fed”
with feed until serious and dangerous out-of-control conditions are reached.
There is also no reason to belive that it is advantageous to operate a kiln
continuously with output rates that are considerably below the rates that
are safely attainable. Each plant should have a set of guidelines that
specify for the operators when kiln production rate should be changed.
Decisions for feed-rate changes should never be based on spur-of-the-
moment beliefs but should instead be well thought out and planned in
advance. In some plants, the decision for a kiln output change is made
solely by production managers or by the kiln operator. Chances for suc-
cessfully achieving higher output rates are better when both the operator
and the production manager, through dialogue, agree to a warranted feed-
rate increase. In short, it is good practice to obtain the input from both
parties before a final decision is made.
FEED-RATE CONTROL

Based on the author’s operating experience, a kiln is “ready” to produce


more clinker when each and every one of the following requirements are
met
1. All related kiln and grinding equipment is able to handle the
increase in feed rate and output rate.
2. The kiln has operated in a stable fashion for at least 4 h prior to
the intended feed increase. Burning-zone and back-end
temperatures have undergone very little change.
3. The oxygen reading, during the previous 4 h, has been
consistently above 1.3 percent.
4. The black feed in the burning zone, during the previous 4 h, has
consistently been located behind or directly under the end of the
flame.

When these four basic requirements are all met, the kiln in essence is
indicating that more production could be achieved without upsetting the
delicate stability of its operation.
Naturally, any feed increase should be done in a series of small steps
instead of all at once. After each measured increase one whould wait for at
least another 4 h to make sure that the above conditions are met before
another increase is made. It is also important not to make this feed in-
crease less than 2 h before the shift change so as not to create difficult
conditions for the next kiln operator.
Regardless of whether more or less output is required, it is necessary to
evaluate what effect the proposed change will have on the overall prospects
of kiln conditions, fuel efficiency, clinker size, bumability, and adjust-
ments that may have to be made to fuel rate, I.D. fan speed, kiln speed,
and other variables. In general, the production rate can be changed either
by maintianing the same feed-to-kiln-speed ratio and changing the kiln
speed, or by continuing the same kiln speed and changing the feed ratio.
Reactions in the burning zone are different from those in the back end,
depending on which of the two procedures is used to bring about the
change in kiln output rate.
A chart similar to Fig. 20.1, which is a portion of a chart that was
designed for one particular kiln and is applicable to that kiln only, will be
of value to the operator in making these adjustments. Such a chart can be
made for any kiln. First, determine the maximum speed at which the kiln
can be operated. Then determine the maximum feed rate in tons per hour
THE ROTARY CEMENT KILN

FEED RATE, TONS PER HOUR

Fig. 20.1 For any operating kiln, a chart can be prepared showing the com-
binations of kiln speed and feed-t&iln speed ratio that will give any selected
production rate. For example, at a kiln speed of 65 rph, a feed ratio of 0.80
gives a feed rate of 74 tons per hour.

for this speed (This information is available in the engineering data for
the kiln under consideration.) The “feed ratio” on the controller is now
arbitrarily set at 1.00 for this speed and feed rate. The relationship is
expressed by the equation:

smax wmaxr2
=- (20-l)
s2 w2

in which
&llilX = maximumkilnspeedinrph
s2 = requiredkihlspeed
W max = maximum feed rate in tons/h at maximum kiln speed and feed
FEED-RATE CONTROL

ratio of 1.00
w, = required feed rate
r2 required feed ratio.
This equaain can be solved for each of the variables:
Laxw2
r-2 = (20-2)
s2wmax

%laxw2
s, = (20-3)
wmaxr2

s2wmaxr2
w, =
n
(20-4)
4MX
Examples: Assume for a certain kiln that
smax=80
W - = 113.6
Determine (a) the feed ratio controller setting when
S2 = 75 and
w,= 100
Using Eq. (20-2)
r2 = 8ox100 =094
75 x 113.6 ’
Determine (b) the kihr speed required when
S2= 95
r2 = 0.90
Using Eq. (20-3) .
s = 80x95 = 74
2
113.6 x 0.90

It is necessary to compute only sufficient values so a family of curves


can be drawn, as in Fig. 20.1. Future operating adjustments can then be
made from the chart
If the feed ratio is changed but kiln speed is held constant, the buming-
zone conditions remain unchanged for 2-3 h, until the different feed rate
arrives in this zone. After a short time, however, the back-end temperature
THE ROTARY CEMENT KILN

starts to fall off making it necessary to adjust the I.D. fan speed and the
fuel rate.
If the feed ratio is held constant and kiln speed is changed, there will be
an immediate reaction in the burning zone and the back end (both cooling
down) that demands an adjustment in I.D. fan speed and fuel rate almost
simultaneously with the changing of the kiln speed.
Neither of the two procedures should be carried out hastily. The smaller
the changes, the less the danger of leading the kiln into an upset. Resting
mods of 4-8 h should be held between a series of feed-rate or kiln-speed
changes in order to make sure that operation of the kiln remains stable.
The above holds true only for production increases or decreases; this by no
means applies to a slowdown period when the kiln has been in an upset.
There are four fundamental rules governing feed rate and feed
advancement that the operator must remember: _
1. Never operate a wet-process kiln longer than 10 mitt, or a dry-
process kiln longer than 15 min, without any feed entering the
kiln.
2. Never permit the back-end gas temperature to exceed the
maximum allowable limit.
3. Never increase the output (or feed rate) at the expense of stable
operating conditions.
4. Never delay a necessary counteractive adjustment until the change
in feed-bed depth is under the flame whenever you know in
advance that such a change will take place.
In the following three examples, an illustration is given of the effects
of kiln-speed and feed-ratio changes on kiln-feed residence and heat-
“soaking” times behind the burning zone. The data given in these tables
apply to a given long dry-process kiln. Similar tables can be developed for
any kiln types by using the appropriate dimensions.
Table 20.1, for example, shows a feed-ratio setting of 0.85, a clinker
output rate of 88 t/h (feed rate 136 t/h) at a kiln speed of 80 rph. At this
kiln speed the feed takes 2.3 h to reach the burning zone and 59.2 tons of
feed is preheated and calcined each hour behind the burning zone.
If the feed ratio were changed to 1.00 (see Table 20.2) this kiln would
have to operate at a speed of 68 rph to obtain the same output of 88 t/h. It
is of interest to the reader to observe that the residence time of the feed
behind the burning zone is now 2.7 h and only 50.3 tons of feed have to
be preheated and calcined within each hour. The third example (Table
20.3) shows a more pronounced change at a feed-ratio setting of 1.24.
FEED-RATE CONTROL

Here, kiln speed 55 will produce approximately the same output but the
residence time is now 3.4 h and only 40 tons of feed have to be heat-treated
behind the burning zone each hour.

TABLE 20.1
KILN PERFORMANCE PARAMETERS
AT VARIOUS KILN SPEEDS
(Feed ratio 0.85)

Travel time Feed rate Feed exposed time


FaCtor 11035 Max. feed 200 Only preheat and
Diameter 14.05 ft Max. kspeed 100 calciniug zone
Slope 0.031 fvft Feed ratio 0.85
Length 425 ft* Factor 1.7
*without burning zone

Kiln Travel time Feed Clinker Feed


speed rate output heated
rph n-h h W tph tph
86 128 2.1 146 94 68.4
83 133 2.2 141 91 63.7
80 138 2.3 136 88 59.2
77 143 2.4 131 84 54.8
74 149 2.5 126 81 50.6
71 155 2.6 121 78 46.6
68 162 2.7 116 75 42.7
65 170 2.8 111 71 39.1
62 178 3.0 105 68 35.5
59 187 3.1 100 65 32.2
56 197 3.3 95 61 29.0
53 208 35 90 58 26.0
50 221 3.7 85 55 23.1
47 235 3.9 80 52 20.4
44 251 42 75 48 17.9
41 269 45 70 45 15.5
38 290 4.8 65 42 13.3
35 315 5.3 60 38 11.3
32 345 5.7 54 35 9.5
29 381 6.3 49 32 7.8
26 424 7.1 44 29 6.2
23 480 8.0 39 25 4.9
20 552 9.2 34 22 3.7
THE ROTARY CEMENT KILN

TABLE 20.2
KILN PERFORMANCE PARAMETERS
AT VARIOUS KILN SPEEDS
(Feed ratio 1.00)

Travel time Feed rate Feed qosed time


Factor. 11035 Max. feed 200 Only preheat and
Diameter 14.05 ft Max. kspeed 100’ calcining zone
Slope 0.031 fvft Feed ratio 1
J-w@ 425 ft* Factor 2
*without burning zone
-
Kiln Travel time Feed Clink.er Feed
speed rate output heated
rph min h tph tph tph
86 128 2.1 172 111 80.4
83 133 2.2 166 107 74.9
80 138 2.3 160 103 69.6
77 143 2.4 154 99 64.5
74 149 2.5 148 95 59.5
71 155 2.6 142 92 54.8
68 162 2.7 136 88 50.3
65 170 2.8 130 84 45.9
62 178 3.0 124 80 41.8
59 187 3.1 118 76 37.9
56 197 3.3 112 72 34.1
53 208 3.5 106 68 30.5
50 221 3.7 100 65 27.2
47 235 3.9 94 61 24.0
44 251 4.2 88 57 21.1
41 269 4.5 82 53 18.3
38 290 4.8 76 49 15.7
35 315 5.3 70 45 13.3
32 345 5.7 64 41 11.1
29 381 6.3 58 37 9.1
26 424 7.1 52 34 7.4
23 480 8.0 46 30 5.8
20 552 9.2 40 26 4.3
FEED-RATE CONTROL

TABLE 20.3
KILN PERFORMANCE PARAMETERS
AT VARIOUS KILN SPEEDS
(Feed ratio 1.24)

Travel time Feed rate Feed exposed time


FaCtor: 11035 Max. feed 200 Only preheat and
Diameter 14.05 ft Max. kspeed 100 calcining mne
Slope 0.031 fvft Feed ratio 1.24
Length 425 ft* . Factor 2.48
*without burning zone

Kiln Travel time Feed Clinker Feed


speed rate output heated
rph min h tph tph tph
86 128 2.1 213 138 99.7
83 133 2.2 206 133 92.9
80 138 2.3 198 128 86.3
77 143 2.4 191 123 79.9
74 149 25 184 118 73.8
71 155 . 2.6 176 114 68.0
68 162 2.7 169 109 62.4
65 170 2.8 161 104 57.0
62 178 3.0 154 99 51.8
59 187 3.1 146 94 46.9
56 197 3.3 139 90 42.3
53 208 3.5 131 85 37.9
50 221 3.7 124 80 33.7
47 235 3.9 117 75 29.8
44 251 4.2 109 70 26.1
41 269 4.5 102 66 22.7
38 290 4.8 94 61 19.5
35 315 5.3 87 56 16.5
32 345 5.7 79 51 13.8
29 381 6.3 72 46 11.3
26 424 7.1 64 42 9.1
23 480 8.0 57 37 7.1
20 552 9.2 50 32 5.4
21.

Kiln Starts and Shutdowns

Until now, only individual control functions applying to steady-state


kiln operations have been discussed. This chapter deals with the funda-
mentals of bringing a kiln on line, shutting a kiln down, and optimizing
the kiln during upset conditions to lead the kiln back to stable operation
after a major operating disturbance (upset).

21.1 KILN STARTS

Starting a kiln from cold state, such as after major refractory repairs,
demands special consideration toward the following items:

a) A complete system check to make sure all main and auxiliary


equipment ate ready for operation
b) Drying of refractory lining
c) Expansion of refractory lining, kiln shell, and tires
d) Timing of the moment when the frost feed is given to the kiln
e) Timing of the burning-zone refractory temperature so that coating
is immediately formed and clinker produced at the moment of
arrival of the fmt feed to the burning zone.

These tasks are made easier when an empty kiln is being started, i.e., a
kiln that contains very little or no feed prior to the start-up. In these
KILN STARTS AND SHUTDOWNS

instances, the kiln-start sequences and measured temperature increases can


be predetermined and preprogrammed. This is not the case for a kiln that
had been shut down under load and has cooled down sufficiently to require a
long kiln warm-up to bring it back to operating temperature. In such
instances, clinker will have to be produced within a short period of time
after the main drive has been activated, thus usually creating too fast a
warm-up period with possible damaging effects on the kiln shell and
refractory lining. It is important to examine the special considerations of
kiln start-up more closely.
a) Complete Systems Check. The most useful and essential means of
ascertaining complete system readiness is a written checklist wherein
operating and maintenance personnel can indicate that each important part
of the kiln equipment has been inspected or test run. Failure to check each
system and starting a kiln under the assumption that all units are ready to
go or worse yet, firing-up a kiln in the absence of preestablished schedules,
is bound to produce erratic starts and possibly lead to premature shutdowns
before the kiln is able to reach operating temperatures. Such complete
system checkouts are absolutely essential when the kiln has been shut
down for a long period of time (weeks/months) or when major main-
tenance overhauls have been made.
Each kiln system is unique and different in its design and thus has its
own multitude of items that should be included in the aforementioned
checklist. Hence, not much benefit could be derived from compiling a list
of items in this book. that must be checked when those items would not
apply to all kilns.
The general format of such a checklist, shown in Table 21.1, could be
used by the reader to develop a list that would include all the applicable
kiln equipment under his control.
b) Drying of Refractory Lining. Special attention must be given to
the drying time when new monolithic (cast) refractory linings have been
installed. These linings have to be slowly and completely dried before the
temperature of the lining can be appreciably raised. But, this attention
toward proper drying should not be limited to monolithic linings alone but
should include all refractories that contain appreciable amounts of
moisture. Such moisture might have been acquired during storage or
during installation, when the wet-laying method (bricks installed with
mortar) was used. Finally, there is the ever-present possibility that the
lining might have acquired moisture due to condensation in the kiln
interior or for accidental reasons during the shutdown (e.g., someone forgot
TABLE 21.1
KILN START CHECKLIST
hrcmosed format1
KILN Sl’ARTS AND SHUTDOWNS

to shut off the water to the back-end water spray or oxygen-analyzer tube).
The drying of the lining is a separate operation and precedes the normal
start-up schedule. Drying can take as long as 5 days on relined preheater
and precalciner cyclones or can be as short as 4 h in cases where only a
short section of lining has been replaced. This drying is usually
accomplished by means of an auxiliary gas or oil burner with a relatively
small fire. Burning of a scrap wood pile in the burning zone is still being
practiced in many plants to achieve the same results. During the entire
designated drying period the kiln back-end temperature is usually held
below 400 F (204 C) and firing is done predominantly with the kiln’s
natural draft only (I.D. fan down). Note of caution: Here too, the rules for
complete combustion (the right amount of air and sufficient heat for
ignition of the fuel) apply at all times. Many operators test for proper
kiln draft by holding a rag in front of the burner-hood viewing ports prior
to lighting the drying fire. This is a crude but very effective method for
preventing lighting up with excessive or insufficient kiln draft.
c) Expansion and Volume Changes. Each type of refractory exhibits
different linear-expansion and volume changes during the time it is brought
to operating temperature. Such differences can even exist within individual
refractory classes such as basic bricks. This is the reason why there are
such widespread differences of opinion on the question of how fast or how
slow a kiln should be brought to operating temperature. What complicates
matters further is the fact that most kiln linings consist of five or more
types of refractory, each undergoing different expansion and volume
changes and thus, in theory, requiring their own optimum heating
schedule. A slow heating schedule that might be beneficial for the basic
lining in the burning zone might, at the same time, be destructive to the
lining in the upper part of the kiln. Since basic refractories undergo the
most and fireclay the least changes during the heating period, it makes
sense to develop a heating schedule that predominantly takes into
consideration the properties of the burning-zone lining.
No attempt is made herein to discuss the initial heating of a new kiln
since such starts demand longer periods for heating the kiln and are gov-
erned by the manufacturer’s specifications. Kiln and refractory manufac-
turers usually supervise and specify heating schedules for the initial start-
up of a new kiln. For these reasons, subsequent discussions will center on
the topic of starting a kiln after refractory linings have been replaced.
Several refractory manufacturers recommend that a basic lining should
be rapidly heated in the first hours at a rate of 100-150 C/h (212-300 F/h)
THE ROTARY CEMENT KILN

until 350 C (662 F) is reached followed by a holding period of 2-4 h at


this temperature. This is to prevent the possible formation of the
destructive hydration products Ca(OH)2 and Mg(OH)z. After this holding
period the temperature is usually slowly raised at a rate of 30-60 C/h
(50-100 F/h) until operating temperatures are reached There is widespread
agreement that the temperature should not be raised faster than a maximum
100 C/h (180 F/h) after the aforementioned holding period to prevent
thermal spalling.
During a start-up, an operator must not only concern himself with the
proper expansion of the lining but must also keep close attention to the
expansion of the kiln shell and tires since both of these are located at a
different distance from the heat source and thus exhibit different expansion
behaviors. The important procedure is to allow both to expand at the
proper rate such that there can be no constriction of the kiln shell by the
slower expanding tires. Extra time spent in closely monitoring tire clear-
ance during a start-up period can prevent the all too frequently observed
damages of kiln shells near the tires. Plants, where tire clearance during
normal operation is already close to minimal levels should, as a pre-
cautionary measure, adopt a somewhat slower heating schedule to prevent
any possible damage to the kiln shell.
Periodic turning of the kiln during the heating period is an absolute
necessity to prevent an uneven heating of the kiln’s circumference, i.e.,
prevent “one-sided” expansion of the lining and kiln shell. Linings with a
low coefficient of thermal expansion (fireclay and insulating bricks) pose
the distinct possibility that they could become loose during the first half of
the start-up period hence too frequent turning in the first few hours should
be avoided
d) Timing of the Feed Introduction. Feed should not be given to the
kiln until the back-end temperature is at a level that will allow the feed to
undergo the normal reactions such as evaporation (wet kiln), preheating,
and calcining. With very few exceptions, kiln-feed introduction begins
when this temperature is between 1740 C (30-75 F) below the target
normal-operating temperature. The feed ratio (feed rate vs. kiln speed) is
generally set at 80-90% of desired full production rate to secure a smooth
and steady kiln start Full production rate is normally not attempted until
after the kiln has stabilized at full speed at the initial lower feed-ratio
setting.
It is not necessary to fust reach the operating target in back-end tem-
perature before feed is given to the kiln because kiln speed and feed rates
KILN STARTS AND SHUTDOWNS

are deliberately set lower during the final stages of a kiln start
e) Timing of the Burning-Zone Temperature. Ideally, one would like
to have the burning zone reach clinkering temperature at the precise
moment when the fresh feed arrives in this zone. This is to secure
immediate formation of coating and prevent unburned “raw” feed from
entering the cooler. Unfortunately these ideal conditions are frequently not
met because such timing takes skill and is often a hit-and-miss situation.
Between the time the feed is given to the kiln and the time it arrives in the
burning xone (considered residence time), an operator should make every
effort and closely monitor the kiln to try to achieve this ideal condition for
this is the key to any smooth kiln start and produces stable kiln operations
thereafter.
Figs. 21.1-21.3 are examples of kiln-start schedules that are typical for
the types of kilns indicated. The reader can use these sample schedules, by
making the necessary modifications, to develop his own schedule suitable
for the kiln under his control. Kiln-start schedules should all have guide-
lines for both the burning-zone and back-end temperatures. A review of
past kiln-start performance charts can also shed some light on the question
of what the fuel rates at any given time are and when the main fire should
be lit in the kiln. Such information, if available, should definitively be
part of any kiln-start schedule.

21.2 KILN STOPS AND SHUTDOWNS


There are many reasons why kilns periodically have to be shut down
but unquestionably the most frequent reasons are:

a) refractory failure,
b ) feed-system failure or feed shortages, and
c ) clinker cooler-component failures.

These three conditons, in most cases, demand a decision from the operator
for an immediate unscheduled shutdown of the kiln system There is often
not enough time available in such situations to weigh or discuss the
merits of a shutdown because a few minutes delay could lead to major
equipment damage. Clearly, an operator should be fully trained in what to
do when such situations occur.
Mistakes have been made by inexperienced operators in such stress
situations that have led direclty to costly damages to the equipment. On
F c
I -----_----_---______________________I___~-----*-------------------------------------------~---------------------------
1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36
--_-_. ___

5I2
r
BURNING ZONE TEUPERAl FUF II
2600 1421 ,
2400 1316
2200 1204
2000 1093
1800 982
1600 871 ‘r t

--
1400 760 e 11oa
1200 649
1000 538
800 427

i.-
600 316
400 204
200 93
.-1 _-- ---- .
-- .
_-
600 316
7

I_-
550 288
500 260 . . . I
450 232
400 204
350 177
300 I49 ‘r
250 121 c
200 93
150

1._
66
100 38
50
-----.
10
_-- t
---. t
_-- t
.-- . .
-- -_-
Ws50 55 Wo61 63 65 66 67 68 69 70
.92 .94
___-----___-_---____------------------------------------------------------------------------------------------------~-“-~---~-
.9'5 .96 .97
HOURS FROU START: 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36

Fig. 21.1 Kiln-start schedule (example). Type of kiln: Wet process < 3.6 m (12 A) diameter; Targets: Back-end temperature: 260 C
(500 FL kiln speed: 70 rph, feed ratio: 0.97
P I -_-_-_.
C 1 2 3 4 5 6 7 S 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32
I -----*--------------------------*-----------------------------~----------------------------------------------, 33 34 35 36 1
-----, --- ___
BURNING ZONE TI PERATURB
2600 1427 , . I I I I I I I II
2400 1316 I7 c,, HOLD
’ ’
2200 1204
2000 1093
1800 982
1600 871
1400 760
1200 649
1000 538
800 427
600 316
400 204
200 93
-----. ---_--.
KILN EXIT IP.H. INLET) GAS TEMPERATURE
1700 927
1550 843 . I I I . . . I

1400 760 I

1250 677
1100 593
950 510
800 427
650 343
500 260
350 177
200 93
10 t
--.
KILN SPEED 40 60 7 5 85 90 95 98 101 106 108
----_---------------____________I_______--------------------~--------------------------------------------~------*-------------
HOURS FROM START:
FSED RITIO

1 2 3 4 5 6 7 S 9 10 1L 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36
.BS ‘88 .91 .93 .94 .95

KILN TURNING : - - - - -

I/ l/3 l/2 l/2 ‘1/2


e a c h 3 0 min I e a c h 2 0 min I e a c h 1 5 min I e a c h 1 0 nin I
continious
aux.driva - I
.l.......*..**....~l***~.....*......~.~~..~...........~............*....................*..........*....~.~.......................

Fig. 21.2 Kiln-start eched~k (example). lope of kiln: 4 stage preheater 4.4 m (14.5 A) diameter, Targets: Kiln exit-gas temperature:
900 C (1650 F), kiln speed: 108 rph. feed ratio: 0.95
P I c 1 2 3 , 5 6 7 8 9 10 11 12 13 I, 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36
.-_ .____---------__1-1_------------------------------------------------------------------------------------------
BURNING ZONE TSNPERATURE
2600 1421 I I I I I I I I I II
2400 1316
2200 1204
2000 1093
lSO0 982
1600 871
1.00 760
1200 649
1000 538
800 427
600 316
400 204
200 93
-_---_-_------_ .A
CHAIN OUTLET IINTEBNEDIATE) GAS 'I lElrIPERATURE
1650 899 I
1500 816
1350 732
1200 649
1050 566
900 482
750 399
600 316
45.0 232
300 149
150 66
t
KILN SPEED 30 38 46 50 54 58 61 64 67 69 70 71 72 73
FEED RATIO .86 .88 .90 .91 ‘92
____________________----------------------------------------------------------------------------------------------~-~--------.
HOURS FROM START: 1 2 3 4 5 6 7 E 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36
KILN TURNING : -c--L - --I-
113 112 l/2 l/2 CO"tl"lO"s I
each 30 ml" each 20 nun I each 15 mm I each 10 mln,aux. drive
I
:i=iiiiiiiiiilii~~EE~===~=~~*~~~~~**~*=*=~=====~~~~~~-*=-==-- --**=*iiii_ -*lixlill **I==iifSiiifiPl==j)-==iiiiiilliS__ --*.*=.i*..***:

Fig. 21.3 Kiln-start schedule (example). Type of kiln: Dry pmcess > 4.5 m (l’5 ft) diameter; Tar@% Chain outlet temperature:
800 C (1480 9, kiln speed: 74 I#, feed ratio: 0.94
KILN mARTS AND SHUlBOWNs

the other hand, it must also be mentioned that many operators have saved
their company production delays and costly equipment repairs by keeping
their composure and using common sense on many occasions when the
kiln seemed to get out of control.
Just as during a kiln start, there are several extraordinary conditions an
operator has to be concerned with when shutting a kiln down. These are:
a) Proper cooling schedules to allow the kiln and refractory lining to
contract in a predetermined manner.
b) Proper kiln-turning schedules to allow the cooling to proceed
evenly on the circumference of the kiln.
c) Emptying of the fuel (especially when fling coal), the feed, and
dust-collecting systems prior to the complete shutdown of the
kiln.
d) Emptying of feed from the kiln, situations permitting, and in
some instances shooting the coating before appreciable cooling
has taken place.
e) Protection of the burner pipe, kiln nose castings, and kiln-hood
area from the possible excessively high temperatures.

Here too, written standard operating procedures to cover these items are
of great help to the operator because kiln shutdowns, hopefully, are not
frequent occurrences. As with any extraordinary kiln situation, operators
tend to quickly forget the sequence of procedures they should follow and
must therefore have some written material at their disposal to periodically
review these procedures.
At this point, it is important to examine these items in detail.
a) Thermal contraction of the kiln. A 160 x 4.7 m (525 x 15.5 ft)
rotary kiln contracts approximately 0.4 m (16 in.) in its length and
approximately 12 mm (0.5 in.) in its circumference during cooling.
Likewise, a ring of refractories also contracts during cooling although at a
different rate than steel. Too fast a cooling rate can induce thermal stresses
on both the lining and the kiln components, hence, the rate and manner in
which a kiln is cooled becomes of paramount importance.
Short clinker inventories or tight production schedules often force
operators to limit kiln shutdowns to as few hours as possible. Many
operators think that so-called forced (fast) cooling of the kiln with the help
of the I.D. fan is a simple means to reduce the time requirements for
allowing workers to enter the kiln. In doing so they tend to forget that
THE ROTARY CEMENT KILN

such rapid cooling could produce costly thermal damage to the kiln
system.
Although no standard procedures will apply to all types of kilns, the
author has found that adherence to the following guidelines can minimize
the risk of thermal shock to the equipment.

(1) Kiln-cooling schedules should provide for a gradual cooling of the


burning-zone refractory temperature at a rate of not more than 100
C (180 F) per hour.
(2) Smaller kilns (< 4 m diameter) should be cooled at the above
stated 100 C/h rate whereas larger-diameter kilns should not be
cooled at a rate higher than 75 C/h (135 F/h).
(3) The I.D. fan should be shut down immediately as soon as the fire
is taken off (fuel cut-off) and should not be used for forced cooling
of the kiln in the frost 6 h after the kiln shutdown. This is an
absolute must on dry- and wet-process kilns equipped with
internal heat exchangers such as chains.
(4) When the fire is taken out, I.D. fan inlet dampers (kiln-draft
control dampers) should be closed as much as permissible during
the first 3 h to prevent excessively high, natural kiln draft (natural
draft occurs without the I.D. fan running).

These guidelines will immediately raise the question of the possibility


that positive hood pressures might occur that are also destructive to the
kiln equipment. To counteract such conditions, the operator must imme-
diately reduce the total amount of air given to the clinker cooler com-
partments as soon as the 1.D: fan is shut down and/or the kiln-draft damper
is being closed.
Note of caution: Make sure the kiln draft is not reversed. In other
words, regulate the dampers such that the kiln gases never are drawn from
the burning zone into the cooler and the cooler stack
b) Kiln-turning schedule. To secure uniform cooling, the kiln has to
be turned on a regular schedule because the feed bed and the refractory
underneath it cools much slower than the refractory wall that is exposed to
the kiln gases. An example of such a turning schedule is shown in
Chapter 3. Turning schedules should not be compromised in the first 2 h
of a shutdown. In other words, when the time comes to turn the kiln,
workers should get clear of the trunnions, preheater, and the cooler interior
to allow the operator to proceed with the turn.
KILN STARTS AND SHUTDOWNS

Note: Make sure they are indeed clear before activating the kiln drive.
On any prolonged kiln shutdown an operator must also make
provisions for evacuating the material from the cooler that is being
discharged from the kiln with every turn. One doesn’t want to be
confronted with a situation where the cooler drive can not be started
because of a material overload in the cooler.
c) Emptying of the coal, feed, and dust-collecting systems. Time and
conditions permitting, all these systems should be emptied out before the
fire is taken out of the kiln. This requires a preplanned sequence of events
that have to be factored into a kiln shutdown schedule to account for the
activities covering the period of l-2 h before the fire is taken out. Emer-
gency conditions excluded, a gradual, sequenced reduction of the kiln con-
trol variables prior to “fire-off” is always preferable to a sudden large
change in these controls.
d) Emptying of the feed from the kiln interior. It takes special skill
and experience to “burn out” a kiln before the fire is taken out and the kiln
drive stopped. Emptying a kiln of feed before the actual shutdown
unquestionably makes it easier to work in the kiln interior especially when
refractory repairs have to be done in the upper part of the kiln. ‘It must
however be emphasized that only experienced operators should be allowed
to “bum” a kiln out, because any miscalculation on the part of the operator
could lead directly to overheated conditions, either at the back end or the
cooler. So-called chain and coal-mill fires have occurred as a direct result
of inexperienced operators attempting to burn out a kiln.
The best one can hope for is to bum out all the feed behind the burning
zone. Empyting the burning zone itself while the main fiie is still lit is
far too risky and should never be attempted.
During the entire period of a burnout the operator must adhere to a firm
set of rules and must abort his attempts (i.e., immediately shut the system
down) whenever any one of these rules are not met, The rules are:
(1) Never allow the oxygen in the kiln exit gases to exceed 1%.
(2) Never allow the back-end temperature (I.G.T. in kilns with
chains) to exceed a predetermined maximum limit.
(3) During the entire period of a burnout, the I.D. fan, fuel, kiln
speed, and cooler air flows must be progressively reduced to
maintain a continuous, gradual decline in backend temperatures.

e) Protecting the burner pipe. The hood and the kiln discharge area
THE ROTARY CEMENT KILN

remain at elevated temperatures for several hours after the fire has been
taken out of the kiln. Some kilns ate equipped with movable burner pipes
to allow for partial or full retraction of the burner. Others have to rely on
the primary air fan to provide the necessary cooling of the burner during
the first 5 h after a kiln shutdown. In these cases, it is advisable to leave
the primary-air fan running at reduced speed until such time as the kiln
interior has cooled to full blackness.

21.3 KILN CYCLES

One of the most common kiln upsets is the condition in which the
back-end and burning-zone temperatures deviate periodically and by a large
degree from the optimum range, forcing the operator to execute a kiln
slowdown in regular intervals. The time span of eachcycle usually is
identical to the time it takes for the feed bed to travel from the preheat to
the burning zone. Cycling kiln conditions can extend from four to six
cycles and, in extreme situations, last as long as several days.
Wet- and dry-process kilns with chain systems are especially subjective
to these cycles. Preheater and precalciner kilns very seldom undergo cycles
because in these kilns calcining is being done predominantly outside the
rotary kiln proper and can be much more easily controlled. Kiln cycles are
the result of irregular calcining conditions within the kiln system.
Calcining conditions are changed when:

a) feed hold-back occurs in the chain system


b) the feed bed becomes fluidized in the calcining zone
C) the kiln is being fed at irregular feed and fuel rates, and
d) most important of all, when the operator, under any one of these
conditions, is not capable of controlling the back-end temperature
within the desired operating range.

Fig. 21.4 is a classic example of a cycling kiln and will be used in the
ensuing discussions to explain what takes place during a cycle and how to
work oneself out of one. In Fig. 21.4 the key variables and controls that
govern the kiln cycles can be recognized. They are: BZT (burning-zone
temperature), and BET (back-end temperature), both of which ate primarily
controlled by the kiln speed, I.D. fan, and the fuel rate. The prevailing
conditions in this figure are also detailed in Table 21.2.
R Z T

B E T

KILN SPEED

I.D. FAN

F U E L

---_-__------------____________________I---------------------------------

1 2 3 4 5 6 7 8 9 10 1 1 12 13 14 15 16 17 18 19 20 2 1 22 23 24
T I M E I N T E R V A L

Fig. 21.4 Typical behavior of cycling kilns


TABLE 21.2
TYPICAL CONDITIONS DURING KILN CYCLES
(as shown in Fig. 21.3)
Time Kiln speed I.D. fan Fuel rate
interval BZT BET Status Action EMUS Action status Action

1 OK OK llol-llld normal normal


2 OK OK n0l-d normal normal
3 OK OK llOI-tllal normal normal
4 down trend OK normal normal high increase
5 too low OK decrease decrease high decrease
6 up trend too high low increase low n0RXXil
7 too high too low n0lllXil high increase low decrease
8 down trend low normal high decrease low increase
9 low OK normal low high
10 OK low IlOlllXil nor@ normal
11 high low normal high low increase
12 too low OK decrease high decrease high decrease
13 up trend too high low increase low tlOllIld
14 too high low normal high iXXRaSe low decrease
15 down trend too low normal high decrease low increase
KILN STARTS AND SHUTDOWNS

The time intervals in this figure could be expressed in increments of 10,


20, or 30 min depending on the length and diameter of a given kiln but
doesn’t alter the indicated typical cycling behavior of these kiln variables.
Cycles are characterized by a heavy and sudden onrush of feed into the
burning zone (time intervals: 5, 12, 19) forcing the operator to execute a
kiln slowdown. After a given period of operating the kiln at this lower
speed, wherein the visibility in the burning zone is usually extremely
poor, the kiln suddenly and rapidly heats up. Once visibility improves,
the operator is normally confronted with a condition in which the load in
the burning zone is very light and dangerously overheated (time intervals
7, 14,21). This in turn compels the operator to make drastic adjustments
in fuel rate and I.D. fan speed. Under certain conditions it might be
necessary to increase the kiln speed above the normal target to prevent
balling up of the clinker bed in the burning zone.
Such changes are dictated by the prevailing dangerous conditions in
which an operator usually has no other choice than to go to a slowdown
procedure to prevent raw feed from entering and damaging the clinker
cooler. Likewise, afterwards the drastic fuel cuts at intervals of 7, 14, and
21 are necessary to guard against possible overheated conditions in the
burning zone that could melt the coating and damage the refractory.
After a kiln has gone through several of these repeat cycles, an operator
could easily become thoroughly disgusted with the kiln or become resigned
in the belief that such instability is inevitable and that there is not much
that could be done to eliminate the problem There are isolated cases
where this might hold true, such as when the kiln-feed properties, kiln
speeds, and feed rates are not compatible to the design of the kiln and the
chain section. However, such cases are more the exception than the rule.
The majority of these cycles am caused by control actions that are focused
on the immediate problem without due regard for the aftereffects these
actions might have on kiln stability.
In the cited example, the kiln slowdown at time interval 12 is clearly
the result of the low back-end temperature at point 7 combined with the
high burning-zone temperature (and consequent low fuel rate) at point 11.
The solutions to cycling kilns is to first study the recording charts to
determine if there is a clear trend in burning-zone temperature behavior,
i.e., determination should be made as to what future time interval a drastic
change can be expected. The tip-off in this respect is the back-end
temperature at a given previous time interval. Once this time lag and trend
is known, kiln control must become a matter of anticipation, i.e., making
THE ROTARY CEMENT KILN

control adjustments before the key variables (BZT and BET) have a chance
to undergo the cycling behaviors.
Assume that the present kiln operation is represented by point 10 in our
example. Assume also that the kiln has gone through several cycles and a
clear trend has been established. From past trends it should become
apparent that, despite prevailing normal burning-zone temperatures and
normal settings for I.D. fan and fuel rate, a short period of higher BZ
temperatures followed by a “push” can be expected shortly because of the
low BET at point 7. Equally important is the observation that the BET is
trending lower at this point 10. This is then the opportune time for a
change in control strategy to break the cycle.
The author has successfully broken many of these repeat kiln cycles by
employing somewhat contrarian procedures when similar conditions to
those shown in time interval 10 prevailed. It is important that these
preventive adjustments are made BEFORE there is a visible sign of a
drastic change in burning-zone conditions. These control adjustments are
(at point 10):

1. REDUCE THE FEED RATIO TO REDUCE THE OVERALL


KILN OUTPUT UNTIL THE KILN IS AGAIN STABILIZED
(i.e., if during previous cycles the feed ratio was set at 0.92, set
the ratio to 0.86).
2. REDUCE KILN SPEED TO HALF AND EXTEND THIS
SLOWDOWN PERIOD TO DOUBLE THE TIME OF
PREVIOUS SLOWDOWNS.
3. GREATLY REDUCE THE I.D. FAN SPEED AND MAINTAIN
THE PREVAILING LOW FUEL RATES. (Note that one can
expect the BET to automatically go up because of the lower kiln
speed.) Keep a close eye on the oxygen and try to maintain a
slightly hotter-than-normal BZT for the next 7 intervals.
4. FOR THE NEXT SEVEN TIME INTERVALS, MAKE EVERY
EFFORT TO STABILIZE THE BACK-END TEMPERATURE
AND MAKE SURE IT DOESN’T EXCEED THE TARGET
LEVEL (although difficult to attain, it is advantageous to try to
maintain a BET slightly lower than normal).

All four steps are taken simuhaneously. These adjustments are contrary
to normal control procedures but necessary for a successful reestablishment
of stable kiln operations. The feed ratio should not be raised to the
previous levels until the kiln has operated at least 8 h without a kiln
B Z T

B E T

KILN SPEED

I.D. FAN
I I I I
F U E L

I I I IWI I H-F1 l i l l
- - - - - - I- - - I- - - I- - - -I - - -I - - I- - - -I - - -I - - -I - - -I-
- - -I - - I- - -I- - - I- - - -I- - -I- - - I.1
1 2 3 4 5 6 7 8 9 10 11 13 17 lit 15 16 ,- *- Li *> Iv
17 1R
&, ”‘

T I M E I N T E R V A L
Fie. 21.5 Control actions to combat cvclinp I&Q
THE ROTARY CEMENT KILN

slowdown. The aforementioned principles for breaking a kiln cycle are


shown in Fig. 21.5. There is no guarantee that these procedures will work
all the time but they are certainly preferable to “fighting” a cycling kiln for
hours or even days.
It should be remembered as mentioned earlier, that kiln cycles can be
caused by faulty design of chain systems both in wet- and dry-process
kilns. Experience has shown that certain systems tend to retain the feed in
the chains (feed hold-back) when the kiln is operated at slow speeds. Once
the kiln is raised to normal levels, the system will release the accumulated
load which in turn will arrive 1-11/2 h later in the burning zone in form of
a “push.” When there are definite indications that these are the reasons for
cycling kiln conditions it is advisable to keep very low kiln speed periods
to a minimum, i.e., get back to say 45 rph as quickly as possible
22.

The 27 Basic Kiln Conditions

In the previous discussions on burning-zone and back-end temperature


conttol, reference was repeatedly made to the necessity of not operating a
kiln by giving attention to one single kiln variable while allowing the
other variables to arbitrarily move about the recording charts in a random
fashion.
The ultimate aim of efficient, stable kiln operations can only be
attained when an operator continuously strives to achieve steady-state
conditons at both ends of the kiln, i.e., maintaining both the burning-zone
and the back-end temperatures within a narrow range. Making a control
adjustment to alleviate a problem at one end of the kiln and neglecting the
effect this adjustment might have at the other end will almost certainly
lead to upset and unstable conditons later.
This important principle of kiln operation can be emphasized using two
examples. In these examples, “actual” refers to the present kiln situation
and “target” represents the optimum condition, i.e., what the temperatures
should be.

Before reading on, it would be helpful to ask oneself the question, “What
would I do in such a situation?” allowing oneself the time to think it over.
THE ROTARY CEMENT KILN

Real kiln conditions are usually not as simple as those depicted in these
examples, but taking herein a simplified approach is helpful in illustrating
the important aspects discussed in this chapter. Assuming that there is no
major imminent upset in the making one should formulate a course of
action and then read on.
Clearly what prevails in this example is a low burning zone, a high
back-end temperature, and moderately high oxygen content, If your answer
was: “Increase the fuel,” you clearly forgot to take the back end into
consideration. Increasing the fuel in such a situation would correct the
burning-zone temperature but would also drive the back-end temperature
higher which, in this example, is definitely not desirable for the sake of
kiln stability. The answer should be: “Decrease I.D. fan.” This single
adjustment would correct three deviations from target at the same time.
Lowering the I.D. fan speed draws less heat to the rear of the kiln and thus
increases the burning-zone temperature, decreases the back end, and also
decreases the oxygen content The answer: “Increase fuel and decrease I.D.
fan,” is also acceptable in this example provided that these actions don’t
drive the oxygen too low and start to show combustibles.
Example 2 is presented without explanation and can be used by the
reader to test his grasp of the fundamental principles discussed above.

Example 2. Actual Target


F C F’ C

Burning-zone temperature 2740 1504 2700 1482


Back-end temperature 1470 799 1500 816
, OXYEW (o/o) 0.6 1.0-1s i
The examples shown above have been used to demonstrate to the reader
the basic, simplified thinking process that must go through an operator’s
mind before making an adjustment It is always necessary to check:

1. the burning-zone temperature,


2. the back-end temperature,
3 . the percent oxygen in the kiln exit gases,

and then and only then, make a decision about which of the controls (fuel,
ID. fan, and/or kiln speed) needs adjustment. This principle of control
applies to any kiln, whether it is a wet, dry or preheater kiln. It even
applies to precalciner kilns, but since they have two firing locations, they
THE 27 BASIC KILN CONDlTlONS

require application of this principle to both parts separately, namely the


rotary kiln and the preheater tower.
Most kiln conditions are not quite as simple as depicted in the above
two examples. There are many more complex problems an operator can be
confronted with that will test his special skills. A more complex situation
is shown in Example 3:

Example 3. Actual Target


F C F c
Burning-zone temperature 2750 1510 2700 1482
Back-end temperature 1600 871 1500 816
Oxygen (o/o) 0.7 1.0-1.5

It is readily apparent that the back-end and burning-zone temperatures


are too high, and the percentage of oxygen is below its permissible mini-
mum The speed of the I.D. fan cannot be reduced as this would result in
an even lower oxygen content and most likely would lead to incomplete
combustion of the fuel. However, one favorable factor is that the burning-
zone temperature is higher than desired. This is important, because a well-
heated burning zone in this instance is an absolute requirement in order for
corrective measures to be carried out. The first step is to reduce the fuel
rate slightly which will increase the oxygen. The burning zone will now
cool down if no further steps are taken. Therefore, as soon as the oxygen
has increased, reduce the ID. fan speed in such a way that the oxygen will
again return to the previous 0.7%. The back-end temperature will now
decrease and theoretically the same burning-zone temperature will be
maintained because with lower I.D. fan speed less heat is transferred to the
rear.
In order to reduce the back-end temperature 100 degrees, the above
described procedure has to be repeated several times in small steps, making
sure to allow for resting periods in order that the reactions can be properly
observed and overcontrolling avoided. The burning zone must remain well
heated throughout the whole procedure. Whenever a cooling down of the
burning zone takes place, the procedure has to be stopped until the burning
zone has warmed up again.
The example is based on the assumption that the kiln has been operat-
ing at full speed for several hours before any attempt is made to lower the
back-end temperature. Should the kiln speed be lower than full operating
speed, this procedure will not be suitable because any kiln speed increase
THE ROTARY CEMENT KILN

would automatically lower the back-end temperature. This example has


been given for the sole purpose of showing how the back-end temperature
can be lowered even when the kiln has been operating under fairly stable
conditions for many hours or even days. It is a fact that many rotary kilns
are operating unnoticed under excessivley high back-end temperatures.
Low oxygen contents do not necessarily mean that one should attempt to
operate the kiln with high Id fan speed and fuel rate.
Now consider another example in which the back-end temperature is
1400 F (760 C), burning-zone temperature is 2800 F (1540 C) and oxygen
is 1.5%. In this case again assume that the kiln in question is operating at
full speed. The burning-zone temperature and the oxygen both are at the
desired level but the back-end temperature is lower than required. First
increase the I.D. fan speed slightly. This will cause the back-end tempera-
ture and the oxygen to increase. The burning-zone temperature will now
decrease if no further steps are taken. As soon as the oxygen indicator
shows a higher reading, increase the fuel rate in such a fashion as to allow
the oxygen to again return to the previous level of 1.5%. This fuel
increase will compensate for the heat removed to the rear of the kiln by the
I.D. fan-speed increase.
Here again one has to make sure that the procedure is undertaken in
small steps and if any change in the burning-zone condition is observed,
the procedure has to be stopped until the burning zone has again returned
to normal.
In this chapter, the terms ideal and proper back-end temperature have
been used, which raises the question: “What is the correct back-end temper-
ature and how can this correct temperature be determined?” This cannot be
readily answered because many factors such as feed rate, kiln speed, feed
composition, moistum content of the feed, and others, all have an in-
fluence on this back-end temperature. However, certain conditions are
known to be the direct result of incorrect back-end temperature. The
following are signs that the back-end temperature is too high:

a) Kiln is operating continuously with a high oxygen content in the


exit gases.
b) Continuous easy-burning conditions are identified by the con-
dition in which clinker formation starts too far back from the
flame. The burning zone is unusually long.
c) Past kiln performances reveal that the kiln has been operating
THE 27 BASIC KILN CONDITIONS

under stable conditions previously for prolonged times with a


lower back-end temperature, under almost identical conditons of
kiln speed, feed rate, feed composition, etc.
d) Kiln is operating at poor fuel efficiency, more fuel being used
than would normally be requried to bum the same amount of a
given feed composition.
e) There is lower than normal feed moisture of the sample taken
from the downstream end of the chains in a wet-process kiln.

A back-end temperature that is too low might be indicated by the


following:

a) Kiln operates for an extended period of time at low oxygen


content in the exit gases.
b) Difficult burning conditions for prolonged time caused by par-
tially calcined feed entering the burning zone.
c) Higher than normal moisture content of the feed sample taken
from the downstream end of the chains in a wet-process kiln,

It is important that the kiln is in a relatively stable conditon before


attempting an investigation into the back-end temperature conditions.
Studies of past kiln performance by means of kiln logs and recorder charts
will reveal valuable information as to whether a kiln is operating at set-
tings that are very close to previous, stable long-time runs.
The reader is reminded that a close control over the back-end temperature
is essential to maintain stable kiln conditions. Chances of obtaining
stable kiln conditions are practically nil if the back-end temperature is
allowed to fluctuate freely.
By now the reader must realize that a kiln cannot be expected to operate
indefinitely in a state of equilibrium. Fluctuations do occur, regardless of
whether a kiln is manually or automatically controlled, and it is the
operator’s responsibility to see that the proper adjustments are ma&
whenever conditions in the kiln vary outside the range of acceptability.
Although there are many variables, with a multitude of combinations, it
has been found that there are three key variables that are of prime
importance. Except under emergency or upset conditons, the operator will
find that these three variables can be maintained within reasonable limits
by means of adjustments to one or more of the three basic controls.
THE ROTARY CEMENT KILN

22.1 THE THREE BASIC VARIABLES AND


CONTROLS

Conditions in the kiln are indicated by:

a) The burning-zone temperature, which is the dominant influence


on the quality of the ptoduct.,
b) The back-end temperature, which is the principal control on
operational stability, and
c) The percentage of oxygen in the exit gas, which governs
combustion conditions and fuel efficiency.

Because any of these conditions can be within the allowable range,


below the minimum allowable value, or above the maximum value, there
are 27 possible basic conditions that will be encountered by the kiln
operator. These are shown graphically in Fig. 22.1. Note that a case
number has been assigned to each condition as an aid to identification.

Fig. 22.1 Evaluation of the three basic variables, burning-zone temperature


@Z), percentage of oxygen in the exit gas (OX), and back-end temperature (BE),
identifies the case number which can be found in Table 22.1 to assist the
operator in controlling his kiln.
THE 27 BASIC KILN CONDITIONS

Basic Controls. In most conditions, except for the emergencies and


upsets listed, control of the three basic variables can be effected by
adjustments to the rate of admission of fuel to the burner, changes in the
speed of the kiln, and changes in the speed of the I.D. fan.
Application Of Control Procedures. First, it is necessary to establish
targets for the three variables: temperature of the burning zone, back-end
temperature, and percentage of oxygen in the exit gas. This information
will be supplied to the operator by his supervisor, and depends on the
composition of the raw feed, type of clinker being burned, and other
factors. Similarly, ranges are established through which these variables
can safely be permitted to fluctuate, because it is obvious that any variable
cannot be held exactly on target at all times. Once these targets and ranges
have been established, then Fig. 22.1 and Table 22.1 can be used to aid in
determining the proper procedure to correct an out-of-control condition.
Abbreviations used in the table and figure are: BE, back-end temperature;
BZ, burning-zone temperature, and OX, percentage of oxygen in kiln exit
gas. Of course, good judgment is necessary in evaluating the conditions,
making sure that no upset or emergency conditions exist.
Example: Assume that in a certain kiln the following values have
been established:

~1

Then BZ is low if below 2750; OK at 2750-2850; high if above 2850.


BE is low if below 1430; OK at 1430-1470; high if above 1470; OX is
low if below 0.4; OK at 0.4-2.0; high if above 2.0.
At the time under consideration, assume that burning-zone temperature
is 2650 (BZ is low), percentage of oxygen is 2.8 (OX is high), and back-
end temperature is 1500 (BE is high). In Fig. 22.1 following the line BZ
low, OX high, BE high, conditions axe found in case No. 9. Under case
No. 9 in Table 22.1 the correct action is listed.
A similar procedure is followed for correcting any of the 27 listed
conditions. The operator should be alert to remedy any out-of-control
condition as soon as the limits of tolerance set up for the particular kiln
and materials are reached. Corrections should be prompt, but care must be
exercised to avoid overcontrolling which could possibly lead the kiln into a
THE ROTARY CEMENT KILN

cycling condition.
Emergencies and Upsets. The control procedures described in this
section are not adequate when the kiln is in an upset condition, or an
emergency exists. Special procedures are necessary under such conditions.
Briefly, these conditions are:
a) Rapid formation of a ring
b) Loss of coating or ring
c) Red spots on the kiln shell
d) Balling of clinker in the burning zone (sausaging)
e) Dangerously high back-end temperature
f) Combustibles in the flue gas
g) Loss of kiln feed and uneven feed-bed depth
h) Unburned feed in the cooler
i) Upsets in which the kiln speed is either higher or lower than
normal.
j) Kiln startup or shutdown periods
k) Any mechanical or electrical equipment failure that interrupts in
any way the regular flow of material, gas, fuel, or air.
Whenever any of these conditions exist, it is necessary to apply the
corrective measures recommended in special sections of this book. The
reader should refer to them and become familiar with them.
The preceding discussions of the 27 basic kiln conditions were‘highly
simplified to show the thought process an operator must follow when
controlling a kiln manually. The advent of the inexpensive
microcomputer has made it possible to refine and extend this control
concept wherein other variables that are also instrumental toward kiln
stability can be included. Computers are capable of scanning all these
important variables in a split second and, if properly programmed, can
make the necessary adjustments to either maintain or to lead the kiln into
stable operations.
The author has developed the software for a training program that is
based on the 27 basic kiln-control concept but that also takes into account
such other factors as the specific heat input requirements, effects of
secondary air temperatures, kiln-drive amperage, and time trends in back-
end temperatures. Samples of this control concept are shown in Table
22.2 and Table 22.3.
Table 22.2 is of interest because it represents condition 9 that has been
extensively discussed earlier (Example 1) with the significant difference
that this time the kiln speed is already below normal therefore demanding
different considerations.
TABLE 22.1
KILN OPERATING PROCEDURES

a. Reduce kiln speed To increase burning zone and back-end temperatures


b. Reduce fuel To move the oxygen percentage into range
When BZ is slightly low:
c. Increase I.D. fan speed To raise the back-end temperature and the oxygen percentage for action (d)
d. Increase fuel rate To raise burning-zone tempemture and to return oxygen percentage into

a. Reduce kiln speed To raise burning-zone temperature


b. Reduce fuel rate oxygen percentage for action (c)
ack-end temperature (but be sure the BE is in a definite down-

When BZ is drastically low:


a. Reduce kiln speed To raise both the back-end and burning-zone temperatures
When BZ is slightly low:
c. Increase I.D. fan speed To raise back-end temperature and oxygen percentage for action (4
d. Increase fuel rate To raise burning-zone temperaWe
TABLE 22.1 (cont’d.)

a. Reduce kiln speed


b. Reduce fuel rate
c. Reduce I.D. fan speed
d. Increase fuel rate To raise burning-zone temperature

To raise burning-zone temperature


To raise oxygen percentage for action (c)
To lower back-end temperature
When BZ is slightly low and oxygen in higher part of range:
d. Reduce I.D. fan speed To reduce back-end temperature and raise burning-zone temperature

To raise both back-end and burning-zone temperatures


b. Reduce I.D. fan speed To maintain back-end temperature
When BZ is slightly low:
c. Increase fuel rate To raise burning-zone temperature and lower oxygen percentage. Back-end
temperature should be rising also; if not, continue increasing fuel rate, and
increase I.D., fan speed

b. Reduce I.D. fan speed ture, If oxygen is still available after this
TABLE 22.1 (co&d.)

To raise burning-zone temperature


To lower backend temperature and oxygen percentage
To raise burning-zone temperature and lower percentage of oxygen
When BZ is slightly low:
d. Reduce I.D. fan speed To raise burning-zone temperature and reduce back-end temperature and
oxygen percentage

b. Increse fuel rate To maintain burning-zone temperature

ing-zone temperature
TABLE 22.1 (cont’d.)

ce back-end temperature

a Increase I.D. fan speed To raise back-end tempera&e


b. Increase fuel rate To maintain burning-zone temperature and reduce percentage of oxygen

a. Reduce I.D. fan speed To lower percentage of oxygen


TABLE 22.1 (cont’d.)

a. Increase kiln speed To avoid overheating the burning zone


b. Increase I.D. fan speed To raise back-end temperature and oxygen percentage
c. Reduce fuel rate
When BZ is slightly high:
d. Increase I.D. fan speed temperature and raise both oxygen percentage and

a. Increase kiln speed To avoid overheating


b. Decrease fuel rate To lower burning-zone temperature
c. Increase I.D. fan speed To increase oxygen percentage and maintain back-end temperature
When BZ is slightly high:
d. Reduce fuel rate To lower burning-zone temperature and raise oxygen percentage

a. hcrease kiln speed To lower back-end temperature and avoid overheating


b. Reduce fuel rate To lower burning-zone and back-end temperatures and increase oxygen
percentage
When BZ is slightly high:
C. Decrease fuel rate To lower burning-zone and back-end temperatures and increase oxygen
TABLE 22.1 (coti’d.)
Action to be taken Reason

When BZ is drastically high:


a. Increase kiln speed To avoid overheating
b. Increase I.D. fan speed To raise back-end temperature
c. Reduce fuel rate To lower burning-zone temperature
When BZ is slightly high:
d. Increase I.D. fan speed To raise back& temperature and lower the burning-zone tempemt~?

23. BZ high When BZ is drastically high:


OX OK & Increase kiln speed To avoid overheating
BE OK b. Decrease fuel rate To lower burning-zone temperature
c. Increase I.D. fan speed To maintain back-end temperature
When,BZ is slightly high:
d. Reduce fuel rate To lower burning-zone tempemture*

24. BZ high When BZ is drastically high:


OX OK a. Increase kiln speed To avoid overheating and lower backad temperature
BE high b. Decrease fuel rate To lower burning-z@ temperature
When BZ is slightly high:
c. Decrease fuel rate To lower both burning-zone and back4 temperabut*
TABLE 22.1 (cont’d.)

To avoid overheating
To increase back-end temperature and lower burning-zone temperature

To raise back-end temperahire and lower burning-zone temperature*

To avoid overheating
b. Increase I.D. fan speed To maintain back-end temperature
To reduce burning-zone temperature

OX high a. Increase kiln speed To lower burning-zone and back-end temperatures


BE high b. Decrease fuel rate
When BZ is slightly high:
C. DeCrease fuel rate To lower burning-zone temperature
d. Decrease I.D. fan speed To lower the back-end temperature*

* If the percentage of oxygen inmases during this adjustment, disqptrl it until the temperatures are brought under contrd.
TABLE 22.2
OPERATING CONDITION 3A
Target Operating
setpoint Actual Deviation range

Burning zone 2700 2660 40 2670-2730


Back-end temperature 1500 1525 25 1480-1520
Oxygen (%I 0.3 O-7-2.0

Combustibles (%> 0 0.1 0.1 0


Back-end temperature. 1.5 h ago 1500 1475 -25
Fuel rate O’W 23316 25470 2154
Heat input (B tu!lb clinker) 2063 2205 142 rt5%
Fuel heat value (Btu/lb) 12600 12320 -280 N/A
I.D. fan speed b-w> 525 513 -12 640-660
Kiln drive (amps) 31 26 -5 29-33
Secondary air O-3 1600 1640 40 1520-1680
Kiln speed (rph) 72 72 0 35-80
Feed rate (lbfi) 225000 224850 -150
Clinker output (lb/h) 142405 142310 -95 N/A
TABLE 22.2 (cont’d)

Kiln St&us Condition 3A


*DANGER-COMBUSTIBLES* BZ low
Overfueling kiln BE high
ox low
I.D. fan is on lower speed
Kiln cooling down
BZ temperature in downward trend
Control action
Immediately reduce fuel rate by 2000 lb
Decrease kiln speed by 15 rph
(when combustibles = 0 and OX > 1.0: Decrease I.D. fan by 3 rpm)
TABLE 22.3
OPERATING CONDITION 9A

Combustibles
Backend temperature
Fuel rate
Heat input (Btu/lb clinker) 2063 2561 498 fS%
Fuel heat value (Btuflb) 12600 12320 -280 N/A
I.D. fan speed (rpm) 525 518 -7 640460
Kiln drive (amps) 31 34 3 29-33
Secondaq air 1600 1170 -430 1520-1680
KilnSpeed &) 72 - 61 -11 35-80
Feed rate (lb/h) 225ooo 181320 -43680
clinker output Wh) 142405 114759 -27646 N/A
TABLE 22.3 (cont’d.)

Kiln statuv condition 9A

Control action

Increase kiln speed 3rph


Decrease I.D. fan 2rpm
23.

Kiln Emergency Conditions

Even in the best-designed and controlled kilns the operator is sometimes


confronted with an emergency situation that calls for immediate and drastic
actions. There is usually not enough time available to consult others as to
the course of action. Quick and appropriate decisions are crucial and must
be decisive for the prevention of possible major damage to the kiln equip-
ment In such an emergency, an operator must ignore kiln-stability
maintenance and concentrate on getting full control over the immediately
dangerous condition. Clearly, the most important requirement in such
situations is that the operator maintains a cool composure.
The following discussions are brief and in a notation format in order to
provide the reader with a quick-access reference. An index has been inserted
to make it possible for the reader to readily find the subject matters in this
chapter. The subjects discussed herein are only a limited sample of the
most common emergency conditions that can be encountered on a rotary
kiln and are by no means complete. They can serve as the cornerstone for
a more extended list of standard operating procedures that must be
separately written and prepared for each type of kiln.
No kiln operator training can be considered complete unless the trainee
has a full knowledge of what to do when such emergencies arise. Since
most likely there is not enough time available to research the procedures or
to ask others what to do, the operator’s reaction must come naturally and
KILN EMERGENCY CONDlTIONS

spontaneously.
Management should review the following operating procedures to make
user they are applicable to their own kiln system The adoption of a
procedure must be appropriate to the particular kiln.

INDEX OF EMERGENCY CONDITIONS

See condition
A) Red spot on kiln shell 23.1
B) Unburned, raw feed in clinker cooler 23.2
C) Ring broken loose inside kiln 23.3
D) Burning zone dangerously hot 23.4
E) Sudden rise in feed-end temperatute 23.5
P) Black smoke emission from kiln stack 23.6
G) Distorted flame shape 23.7
I-I) Loss of refractory lining 23.8
I ) Cooler drive or clinker belt stopped 23.9
J ) Red-hot clinker at cooler discharge 23.10
K) Rapid rise in coal-mill temperature 23.11
L) Power failure 23.12
M) Chain “fire” 23.13
N) Heavy rain or thunderstorm 23.14
0) High, positive hood pressures 23.15
THE ROTARY CEMENT KILN

EMERGENCY CONDITION 23.1:


RED SPOT ON KILN SHELL

Indicators:
l By visual observation

l Shell scanner: Sharp and rapid shell-temperature increase to a

level above 450 C (850 F)


l Visual observation of loose refractory bricks in the clinker bed of
the burning zone

Possible Effects and Dangers:


Severe warpage and damage to kiln shell.

Recommended Action to Take:


A) Small red spot, located in upper transition or center of burning
rOtltZ.*

Continue normal operation of kiln but:


1. Start shellcooling fans in the ama of the red spot.
2. Shorten flame to bring “black” feed over area of red spot in an
attempt to form new coating.
3. Maintain normal burning-zone temperatures.
4. Change kiln-feed chemistry to obtain easier-burning mix.
B) Large red spot, located under or near a kiln tire or in areas where
usually no coating is formed:
Shut kiln down immediately
WARNING: Under no circumstances should a water spray be used
on the red spot as this could immediately result in
severe kiln-shell damage.

Possible Measures to Prevent Reoccurrence:


l Make sure flame configuration and characteristics are not causing
localized coating erosion or continuous, excessive overheating.
l Employ proper refractory installation methods.
l Minimize frequency of kiln shutdowns and upsets.
l Avoid “hard”-burning mixes (i.e., ensure sufficient percent liquid
content in mix to promote coating formation).
KILN EMERGENCY CONDITIONS

EMERGENCY CONDITION 23.2:


RAW, UNBURNED FEED IN CLINKER COOLER

Indicators:
l Onrush of raw feed into and beyond burning zone
l “Black feed” position advanced more than halfway under the flame
in burning zone
l “Black-out” in burning zone
l Red grates in cooler
l Rapid rise in cooler grate and clinker discharge temperatures
l Cooler drag-chain amperage increases rapidly

Possible Effects and Dangers: I


l Thermal damage to cooler grates and grate-drive mechanism
l Flame extinguishment in burning zone
l Fire on clinker conveyor belts
l Excessively high temperatures in coal-mill air circuit
WARNING: Watch for incomplete combustion when visibility in
burning zone is severely restricted.

Action to Take: First and foremost-don’t wait until raw feed is in


the cooler; act when the first signs of impending problems are visible
in the burning zone.
1. Immediately reduce kiln speed to minimum (or turn kiln on
auxiliary drive).
2. Reduce fuel and I.D. fan speed in accordance with standard
slowdown procedures to protect kiln back end.
3. Reduce cooler-grate drive speed (switch to manual control) to
allow material in cooler more time for cooling.
4. Adjust cooler air-flow rates to obtain maximum cooling without
the hood pressure going positive.
5. Advise all unauthorized personnel to stay clear of the firing floor,
cooler, and coal-mill area.
Preventive Measures for Reoccurence:
l Accelerate frequency of visual observation of burning zone for

early detection of impending cooler upsets.


l Evaluate kiln output rates vs. cooler capabilities and kiln-
operating stability.
THE ROTARY CEMENT KILN

EMERGENCY CONDITION 23.3:


LARGE RING BROKEN LOOSE IN KILN

Indicators:
l Visual observation of large junks in burning zone

* Sudden drop in kiln backend draft


* Large drop in oxygen content of kiln exit gases
l Hood pressure tending toward positive side

1 l Sudden change in kiln-drive amperage

Possible Effects and Dangers:


l Overloading cooler with unburned feed

l Onrush of excessive amounts of feed into the burning zone

* Damage to cooler drive and grates


l Large pieces jamming cooler hammer mill

l Red-hot clinker leaving cooler

Action to Take:
When amount of feed and ring fragments in burning zone is
extremely large:
1. Immediately reduce kiln speed to minimum.
2 . Reduce fuel and I.D. fan speed to keep back-end temperature under
control.
3. Switch cooler grate control to manual and reduce grate speed.
4. Adjust cooler air flows to maximum flow possible; without that
the hood pressure goes positive.
5. Have personnel on standby to watch the cooler and the
hammermill for possible overloading, overheating, and jamming.

Possible Preventive Measures for Reoccurence:


l Laboratory to reevaluate chemistry of kiln feed (including dust-

return rates) for possible elimination of ring formation. If no


solution in this area possible then:
l Initiate regular schedule to remove rings and heavy buildup by
means of special &vices designed for this purpose.
l Initiate regular procedures to displace the burning-zone location
on a daily basis (i.e., reposition burner every morning).
KILN EMERGENCY CONDITIONS

EMERGENCY CONDITION 23.4:


BURNING ZONE DANGEROUSLY HOT I
Indicatys:
l Clinker balling (“sausage”) in burning zone

l Coating dripping off the wall

. Sliding molten clinker bed in burning zone


l Burning-zone temperature recording too high

* Cooler undergrate pressures too high


l Yellow/white burning zone

Possible Effects and Dangers:


l Loss of coating and thermal damage to refractory

l Red spots on the kiln shell

l Thermal damage to cooler and kiln-hood components

Possible Actions:
1, Reduce fuel rate to minimum until sausaging stops.
2, Increase kiln speed approximately 5-10 rph until sausage is
broken
3. Provide maximum possible air in cooler (without hood pressure
going positive).
4. Reduce primary air flow.
THEN, AS SOON AS THE PRIMARY OBJECTWE OF
BREAKING THE AGGLOMERATION IS ACCOMPLISHED:
5. Reduce the kiln and I.D. speed AND increase fuel rate to restore
normal operating conditions.

Preventive Measures:
l If “sausaging” is frequent and the result of easy-burning mix, have

laboratory evaluate possibility of providing a mix with less per-


cent liquid content.
o Make more frequent, vigilant observation of burning-zone condi-
tions.
l Evaluate flame position and shape to determine if thinner, longer

flame is feasible.
THE ROTARY CEMENT KILN

EMERGENCY CONDITION 23.5:


SUDDEN, SHARP RISE IN BACK-END
TEMPERATURE

Possible Reasons:
l Feed shortage

l Combustibles in exit gas


l 1.d. fan speed too high

* Ki~speedtoolow
l (&in ‘y-ii”

Possible Effects and Dangers:


l Chainfireonwetanddrykihrs
l Thermal damage to back end, dust collector, and preheater tower

equipment
l Delayed ignition of fuel in back end of the kiln

Possible Actions:
1. Immediately de-energize electrostatic precipitator.
2. Immediately reduce fuel rate and I.D. fan speed to obtain less than
0.3% oxygen in exit gas.
WARNING: Do not cut @fuel rate completely as this could trigger
an explosion.
3. Increase kiln speed and feed rate.
4. Warn personnel to stay clear of kiln back end.
5. Do not open any doors in kiln back end.
THEN, AS SOON AS THE PRIMARY OBJECTIVE OF
BRINGING THE BACK-END TEMPERATURE UNDER
CONTROL IS ACCOMPLISHED:
6. Return kiln control variables to normal to restore normal
operating conditions.
7. Check out back end to determine if thermal damage has occurred,
Preventive Measures:
l Do not operate kiln without feed for more than 10 min.

l Provide alarms and properly maintain kiln instrumentation to


obtain warnings before the back-end temperature gets out of hand.
l Maintain close vigilance over combustion, back end, and feed-
flow conditions during kiln starts, shutdowns, and upsets.
KILN EMERGENCY CONDITIONS

EMERGENCY CONDITION 23.6:


BLACK SMOKE EMISSION FROM KILN STACK

Indicators:
l Combustibles in exit gases
l Oxygen iu exit gas too low
l Blame extinguished for poor ignition conditions
l Burning-temperature too low
l Excessive fuel rates and/or insufficient kiln draft

Possible Effects and Dangers:


l Explosion or thermal damage to kiln back-end equipment

Possible Actions:
1. Immediately de-energize electrostatic precipitator.
2. Immediately reduce fuel rate (don’t shut off) and increase I.D. fan
to obtain:
a ) zero combustibles in exit gas
b ) oxygen between 0.2 and a maximum of 0.5% in exit gas
3. After black smoke has cleared, maintain the low oxygen/zero
combustibles for at least 10 min before restoring kiln variables to
normal.

Preventive Measures:
l Improve control over flame and firing conditions.
l Make frequent, vigilant observation of fuel rates, gas analysis,

flame and kiln-draft conditions during kiln starts and upsets.


THE ROTARY CEMENT KILN

EMERGENCY CONDITION 23.7:


DISTORTED FLAME SHAPE

Indicators:
l Irregular and unusual flame shape
e Fragmented flame where part of flame impinges on lining near
the kiln discharge area.

Possible Effects and Dangers:


* Thermal damage to refractory, kiln shell, and kiln hood
@ Red spots on the kiln shell near discharge area
l Thermal damage to nose castings

Possible Actions:
1. Inspect burner pipe for damage.
2. If flame is erratic and severely impinges upon lining near the kiln
discharge area: SHUT KILN DOWN IMMEDIATELY.
3. If flame is only slightly distort& Adjust burner position and
primary air flow and schedule burner-pipe repairs for next kiln
shutdown.

Preventive Measures:
@ Regular inspection and maintenance of the burner pipe during
each prolonged kiln shutdown.
* Improved protection (castables, air cooling) for burner pipe.
l Maintain primary air flow for at least 2 h after a kiln has been

shut down or pull back burner pipe immediately when kiln is


being shut down.
* Investigate possibility of relocating and redesigning burner pipe
to eliminate frequent damage to burner.
KILN EMERGENCY CONDITIONS

EMERGENCY CONDITION 23.8:


LOSS OF SECTION OF REFRACTORY LINING

Indicators:
* Loose bricks in clinker bed of burning zone
* Delineated (linear instead of round) red spot on the kiln shell
l Rapid rise in shell temperature

Possible Effects and Dangers:


l Thermal damage and distortion of kiln shell and tire
l Further collapse of large sections of linings

Possible Actions:
1. IMMEDIATELY SHUT DOWN KILN.

Preventive Measures:
* Employ proper refractory installation methods and procedures.
0 Make annual checks of kiln alignment and shell ovality.
l Have refractory manufacturer provide uniform shapes and proper
expansion allowance for each type of brick.
* Avoid excessive turning when kiln is cold during shutdowns.
THE ROTARY CEMENT KILN

EMERGENCY CONDITION 23.9:


COOLER DRIVE OR CLINKER BELT STOPPED

Indicators:
0 Cooler overloaded
l Large chunks of coating in cooler
l High undergrate pressures
l High cooler drive amps prior to drive stop
l Clinker transfer chutes plugged

Possible Effects and Dangers:


l Thermal damage to cooler components

Possible Actions:
1. Immediately reduce kiln speed to minimum and attempt to restart
clinker belt and/or cooler drive.
2. If’ drives can not be restarted within 5 min, shut kiln down.
NOTE: After kiln has been shut down, consiak possibility of
turning the kiln in less frequent intervals to prevent further
overloading of cooler (kiln still has to be turned periodically
nevertheless).

Preventive Measures:
l Know at what amperage the cooler drive is likely to fail and

provide alarm for overload


l Adjust kiln parameters (namely kiln speed) before cooler can

become overloaded at the times when heavier feed load is observed


in the burning zone.
I
KILN EMERGENCY CONDITIONS

I EMERGENCY CONDITION 23.10:


RED CLINKER AT COOLER DISCHARGE

Indicators:
I

l High drag-chain amps

l Sudden drop in undergrate pressure (grate out)

l Excessively high undergrate pressure (cooler overloaded)


l Cooler drive amps and clinker bed depth too high

. Cooler loaded with coating and ring fragments


l Stalagmite formation at cooler inlet
l Uneven cross-sectional loading of cooler

* Insufficient air flow into cooler

Possible Effects and Dangers:


0 Thermal damage to cooler components
0 Thermal damage to clinker transport equipment

Possible Actions:
1. Immediately make a visual check of the cooler to determine
reason for red-clinker discharge.
If cooler grate out, SHUT KILN DOWN.
If cooler overloaded, REDUCE KILN SPEED TO MINIMJM
AND REDUCE COOLER-GRATE DRNE SPEED TO ALLOW
MORE TIME FOR COOLING.
2. Increase air flow into cooler.
3. Activate water spray at cooler discharge and reroute clinker to
prevent damage to conveyor belts.

Preventive Measures:
A) On frequent grate failures:
o Investigate for possible faulty grate-installation methods by
maintenance department.
l Investigate quality of grates and bolts used

B) On frequent one-sided loading of cooler bed:


l Investigate possible cooler-design changes.
l Investigate possibilities for elimination of stalagmite
(2nowmen”) formation at cooler inlet.
C) On frequent overloading of cooler due to upsets:
l Slow kiln speed down before raw feed enters cooler or cooler can

become overloaded (make your corrective moves before things get


out of control).
THE ROTARY CEMENT KILN

EMERGENCY CONDITION 23.11:


RAPID RISE OF TEMPERATURE IN COAL SYSTEM

Possible Effects and Dangers:


l Explosion

0 Thermal damage to coal system

Possible Actions:
WARNING: Do not open any doors in the system that could
provide the oxygen for an explosion or more serious
fire.
1. Inject inert gas (COz) into coal-mill inlet.
2. Flood coal mill with kiln feed or excessive coal.
3. Warn all personnel to stay clear of system.
4. Stop or reduce air flow to coal null to minimum.

Preventive Measures for Reoccurence:


l provide coal-mill inlet with magnetic device to extract metal
fragments from coal feeder belt.
l Keep paper, rags, etc. out of coal storage pile.
l Do not feed coal mill with coal that has undergone spontaneous
ignition (“smothering”) while in storage.
l Keep coal-mill detramp chute clear.
l Provide coal-mill system with automatic fire-extinguishing
devices.
l Do not operate coal mill above predetermined safe temperature for
any given type of coal.
KILN EMERGENCY CONDITIONS

EMERGENCY CONDITION 23.12:


POWER FAILURE

Possible Effects and Dangers:


l Warpage of kiln shell

l Thermal damage to burner pipe, instrumentation, and equipment


at kiln discharge area
l On Coal-Fired Kilns: Settlement of ground coal in coal system
that could lead to a fire and/or explosion
I
Possible Actions:
1. Immediately, start auxiliary power generator and primary air fan
(coal-mill fan on direct-fired kilns).
2. Retract burner pipe and protect T.V. monitor in kiln hood.
3. Start quarter turn on kiln not later than 10 min after the power
faih.ue.
4. If mwilubZe. close feed-end damper manually to prevent hot gases
from escaping from kiln by natural draft.
THE ROTARY CEMENT KILN

EMERGENCY CONDITION 23.13:


CHAIN “FIRE”

Indicators:
l Rapid, sudden rise in intermediate- and exit-gas temperatures
l By visual observation

Possible Effects and Dangers:


l Melt-down and loss of chains
l Damage to kiln shell in chain-system area
0 On wet-process kilns: Steam explosion
1 l Thermal damage to kiln back-end equipment

Possible Actions:
WARNING: Under no circtumtances should there be water added at
the feed end.
’ 1. Immediately, reduce fuel rate to minimum (BUT DON’T SHUT
FUEL OFF COMPLETELY! !!). At the same time, reduce ID.
fan speed to obtain zero combustibles and less than 0.3% oxygen.
2. Increase kiln speed and feed rate to maximum until the back-end
temperature is under control.
3. On wet-process kilns: Clear all personnel from firing floor.

Preventive Measures:
l Avoid operating the kiln for more than 10 min when there is a

feed shortage.
l Establish and enforce maximum permissible operating limits for

intermediate and/or exit-gas temperatures.


KILN EMERGENCY CONDITIONS

EMERGENCY CONDITION 23.14:


HEAVY RAIN OR THUNDERSTORM

Possible Effects and Dangers:


On kilns that are exposed to the elements:
l Loss of coating and collapse of refractory lining
l Thermal damage to kiln shell
l Possibility of a power failure

Possible Actions:
If storm occurs shortly after a kiln shutdown:
1. Jack (turn) kiln more frequently or turn continuously on auxiliary
drive.
2. Start auxiliary power generator in preparation for a possible
power failure.

EMERGENCY CONDITION 23.15:


SUDDEN, HIGH-POSITIVE HOOD PRESSURE

Possible Reasons:
l I.D. fan failure

l Large ring or buildup broken loose inside kiln


l Instrumentation failure of cooler air flow, cooler stack damper, or
l I.D. fan control
l Steam explosion on wet-process kilns

Possible Effects and Dangers:


. All personnel on firing floor is in peril
l Thermal damage to equipment on firing floor and hood
l Danger of back-fire in coal system

Possible Actions:
1. Immediately, clear all personnel from firing floor.
2. Immediately, reduce fuel rate to minimum and increase I.D. fan

3. Reduce cooler air-flow rates into undergrate compartments.


4. Open cooler excess air damper manually.
24.

Safety and Accident Prevention

The previous chapters have concentrated on kiln control techniques and


equipment. To do justice to all aspects of kiln operation, however, a
discussion of safety is necessary in a book of this kind. There are many
situations in which a worker could be injured because of a lack of machine
guards, failure to wear proper protective clothing, or faulty job perfor-
mance by himself or another person. A kiln operator must familiarize
himself with all the potential hazards that might exist in and around the
kilns under his control, and set for himself a high standard of safety
consciousness. He especially should be alert to point out hazards to other
employees and should see to it that no employee works in an unsafe
manner on his kilns.
Before entering into a detailed discussion of the hazards around kilns,
fundamentals of safety in general should be reviewed so the reader can
relate them to the rotary kiln.

24.1 SAFETY

s’imply stated, safety measures are introduced into a plant for two
reasons: to protect an employee from injury, either physically or finan-
cially, while performing his work, and to prevent financial loss to the
employer as a result of damage to the equipment or compensation
payments which are a part of nearly every industrial accident. Management
and employee alike are responsible for making the plant a safe place in
SAFETY AND ACCIDENT PREVENTION

which to work, and achieving injury-free work performance day after day.
A plant safety program can only be successful when all parties
wholeheartedly believe in safety, and when safety becomes a part of the
working life of every man in the plant regardless of his position. Evasion
of safety responsibilities by the individual, implicitly delegating such
responsibility to others, generally referred to as “passing the buck,” is
bound to result in failure of the program.
If a supervisor appears to be strict and unyielding with respect to safety
rules and procedures, his efforts should be appreciated, and not resented.
After all, it is the responsibility of the supervisor to see that the employee
first endeavors to make himself a safe worker, and only after he has
accomplished this and is a good example to others can the employee then
try to win others over to the side of safety. That’s what safety is all
about, It is first of all a state of mind, an idea implemented by a
constructive attitude that causes a man to recognize dangerous situations
before an accident occurs. It is not something to be lived with under
duress because it has been imposed in the form of rules by management.
Most importantly, it deserves the support of all employees in the plant

24.2 ACCIDENTS

Now consider the word “accident.” An accident to many workers


represents a condition in which someone is injured and property is
damaged. Anything less is looked on as a close call, a near miss, or a bit
of good luck. To put it in the proper perspective, an accident is an acci-
dent even though no one is injured or no damage is done, An accident is
any unintentional or unexpected interruption of the orderly progress of the
work Accidents do not happen. An accident is the result of some unsafe
act or equipment. We have but little control over the severity of injury
once an accident has occurred, but we can control the conditions leading up
to the injury. A statistical analysis of thousands of accidents and injuries
shows that every accident that resulted in major injury (a lost time
accident) was preceded by 29 minor accidents (no lost time and only minor
injury) and by 300 accidents that caused no injury at all. So-called near
misses and close calls are included in the 300. These statistics warn that,
if we have a great number of close calls on the job, sooner or later there
will be a serious injury in the plant. Ironically, one usually finds the
‘I-I-E ROTARY CEMENT KILN

reasons for an injury accident, but seldom wants to take the time to do
something after a close call to prevent these minor accidents from reaching
major proportions.
For every accident that is the result of unsafe conditions there are nine
that were caused by unsafe acts, including those resulting from failure to
recognize unsafe conditions. An employee can easily fall into the habit of
overlooking some basic safety procedures and taking unnecessary chances
when he develops the attitude that because nothing happened the last time,
nothing will happen the next time he does the same thing.
The “Accident Roundtable,” published montly by the Portland Cement
Association, points out that accidents in the vicinity of kilns have a higher
frequency rate than those in other areas of the plant. It is common practice
m cement plants to provide general safety rules that apply to all employees
throughout the plant. There are, however, certain hazards that are unique
to rotary kilns, and it is these dangers that the kiln operator must become
aware of. Table 24.1 is a compilation of kiln hazards and possible action
that can be taken to eliminate or reduce the dangerous condition.

TABLE 24.1
KILN HAZARDOUS CONDITIONS
Hazard Action to eliminate or reduce hazard

Backfii and l Open either one cooler or burner hood door


explosion during before lighting fm in kiln.
kiln light-up l Secure proper draft in kiln before fire is lighted

(very important).
l Do not allow unauthorized persons to stand near

the burner hood during light-up.


l Stay clear of burner hood ports when igniting
the fuel.
l Avoid excessive fuel flow on initial light-up of
the flame.
l Start the primary-air fan before opening the fuel

valve.
l When fmng coal, make sure that no coal-dust

spills are present on firing floor, around coal


feeder. or in the primary-air pipe.
SAFETY AND ACCIDENT PREVENTION

Hazard Action to eliminate or reduce hazard

ktting any kiln ma- l Make sum all persons am clear of kiln equipment
:flimy into lmtion before each unit is started.
luring startup l Sound horn (if available) to signal startup.

*Inspect all circuit breakers before the startup to


make sure that all safety tags and locks have
been removed.
l Make sure all machine guards are in place before
any equipment is started.

Relining the kiln l Construct a proper bridge across the burner hood
with refractory bricks from firing floor to kiln nose.
and material *Inspect coating and remove loose overhangs
before passing underneath.
l Keep all unauthorized personnel out of kiln
intclior.
l Use protective screen when working under loose
refractory and coating, if no alternate procedure
is possible.
l Any employee working inside the kiln should

have positive means, such as locking out the


kiln drive with his own lock, to assure that the
kiln cannot be started while he is inside.
l Have proper posture and steady footing when

lifting bricks or scaling coating.


l Do not work underneath the burner hood bridge

while material is being hauled in and out of the


kiln.
l Do not test run cooler fans when workmen are

inside kiln.
l Do not run I.D. fan when workmen am at kilt
rear or in chain section.
Working near or on l Wear extra protective clothing to guard against

dust-collecting burns from hot dust.


equipment l Wash skin thoroughly with clear water after

contact with alkahne dust.


l Have a second workman as safety man standing

by whenever working under or in bins or hop-


pers cxmtaining material.
.Do not allow workmen to work inside hoppers
without being properly secured on safety lines
and belts.
THE ROTARY CEMENT KILN

Hazard Action to eliminate or reduce hazard

*When working on plugged flue chambers, be


constantly on guard against potential dust
flushes and cave-in of overhanging material.

Shooting clinker .Do not allow any employees other than the gun
ings with industrial crew on the firing floor during ring shooting.
Sun 0 Do not tamper with the ammunition.
*Keep all live ammunition locked up and away
from the firing floor when not in use.
l Permit only experienced and trained persons to

operate the kiln gun.


l Use ear muffs or plugs when ftig gun.

0 Cotton sutffed in the ear is not adequate.


l Clean gun at frequent intervals and do not
attempt to fire an apparently defective gun.
.If kiln has no chain section, keep all persons
away from the kiln back end and rope this area
off before shooting.

Clinker, fuel oil, and l Clean up any spills immediately.


coal dust spills l Provide adequate clean-up cans and facilities for
easy removal of spills.
l Initiate repair action when spills are caused by

leaks that can be repaired.

Gas, fuel oil, coal, l Report any gas odor on the firing floor im-
and steam leaks in mediately to the foreman.
fuel system l Provide for periodic inspection of fuel and steam
lines and system to detect leaks and other defects
as a preventive measure against major breaks ir
the system.

Burner hood l Do not allow anyone to look into the burning


portholes and cooler zone while the kiln is in operation unless ap
doors proved safety equipment for viewing is used.
l Use proper protective clothing when working

near open burner hood and cooler doors while


the kiln is in operation.
l Instruct all persons to stay clear of the portholes
whenever the hood pressure is temporarily or
the positive side.
Appendix

A: THE INTERNATIONAL SYSTEM OF UNITS


@I)

The following is a guide to familiarize the reader with the units,


prefutes, symbols, and formulas used in the International System of Units.
“SI” is the common language in which scientific and technical data are
presented worldwide.

A.1 Base Units


Quantity unit SI symbol
length meter
mass kilOgEUll G
time second S
electric current ampere A
thermodynamic temperature Kelvin K
amount of substance mole mol
luminous intensity candela cd

A.2 Supplementary Units


plane angle radian rad
solid angle StIXXihl sr
THE ROTARY CEMENT KILN
A.3 Derived Units

Qmit>l unit SI symbol Formula


acceleration meter per second2 dS2
angular acceleration radian per second2 Ed/S2
angular velocity radian per second rad/S
area square meter m2
density kilogram per meter3 kg/m3
electric capacitance farad F As/V
electric field strength volt per meter V/m
electric conductance siemens S AIV
electric inductance henry H VslA
electric potential diff. volt V W/A
electric resistance ohm a VIA
electromotive force volt V WIA
energy joule J N+m
entropy joule per kelvin J/K
force newton N kg-m/S2
frequency hertz Hz (cycle)&
illuminance lux lx Mm2
luminance candela per meter2 cd/m2
luminous flux lumen lm cdsr
magnetic field strength ampere per meter a/m
magnetic flux weber wb v*s
magnetic flux density tesla T Wb/m2
magnetomotive force ampere A
power watt W J/s
pressure pascal Pa N/m2
quantity of electricity coulomb C As
quantity of heat joule J N-m
radiant intensity watt per steradian Wlsr
specific heat joule per kg-kelvin J&K
stress pascal Pa N/m2
thermal conductivity watt per m-kelvin WlmK
velocity meter per second m/S
viscosity dynamic pas&second Pas
viscosity, kinematic m2 per second m2/s
voltage volt V W/A
volume* cubic meter ms
wavenumber reciprocal meter (wave)/m
work joule J N-m

In nomd engineering work, where high precision is not required, the use of the
liter as a unit to express volume is acceptable.
B: WEIGHTS AND MEASURES
B.l Weights
1 lb 0.4536 kg 1 kg 2.2046 lb
1 lb 16 oz lg 0.0352739
1 lb 453.59 g lg 0.0022046 pbz
1 lb 444820 dynes 1 dynes 2.248E-06 lb
1 short ton 0.907185 metric ton 1 metric ton 1.102311 short ton
1 short ton 907.2 kg lkg 0.0011023 short ton
1 short ton 2000 lb 1 metric ton 2204.5 lb
1 lb 0.0004536 metric ton 1 metric ton 1000 kg
1 oz 28.3495 z! 1 kg 35.2739 oz
B.2 Linear Measures
1 in‘ = 254 mm lmm = 0.03937 in.
1 in. = 2.54 cm 1 cm = 0.3937 in.
1 in. = 0.0254 m lm = 39‘370079
lft = 0.3048 m lm = 3.2808399 ii
1 ft = 30.479 cm 1 cm = 0.0328095 ft
lft = 300.479 mm 1nUli = 0.003328 ft
lft =: 0.0003048 km lkm i= 3280.8399 ft
1 stat. mi = 1.609 km lkm = 0.621504 stat. mi
1 stat. mi = 0.8684 naut. mi lkm =: 1000 m
1 naut. mile = 1.1515 stat. mi 1 micron = 25.4 w

lm = 1.0930156 yd
B.3 Areas
-7
1 in? =: 6.4516 cm2 1 cm2 = 0.1550003 in.2
1 ft2 = 144 in.2 1 cm2 = 0.001076 ft2
1 ft2 = 0.092903 m2 1 m2 = 10.763915 ft2
1 ft2 = 929.03 cm2 1 hectare = 2.47 acres
1 ft2 =: 2.296E-m acres li3lE = 1 1 9 . 6 yd2
1 rni2 L 640 lkItI2 = 247.10883 acres
1 mi2 = 2.59 s lkd E 0.3861004 mi2
1 acre = 43560 ft2 1Ztl-e =. 100 m2
lam = 4046.8 m2 1 m2 = O.ooo247l a c r e s
1 yd2 =i 0.8361 m2 1 m2 = 1.1960292 yd2
B.4 Volumes

0.0353145 ft3

1flozUSA = 33.813485 fl oz US
2.1133607 pt US ii
1,0566803 qt us zs
2
2
8
x
2
E
5
1 gr/ft3 = 2.288E-06 g/cm3
1 lb/ft3 = 0.0160185 g/cm3
1 lblyd3 * 0.0005933 g/cm3
1 lb/in) = 2.767997 g/cm3 0.3612721 lb/in?
0.062422 lb/ft3
1 lb/gal US =
B.6 Specific Flow Rates (Velocities)
1 ft%nin 0.02832 m%nin 1 m%nin 35.3 10734 ft%nin
1 ft%s 448.83 gal US/min 1 gal US/min 0.002228 ft%
1 ft/rnin 0.508 cm/s 1 cm/s 1.9685039 ftlmin
1 ft/min 0.018288 lkm!h 54.680665 ft/min
1 ft/rnin 0.3048 rldnin lmlmin 3.2808399 ft/rnin
1 ft/rnin 0.011364 miih lmiih 87.997184 ft/min
1 gal/min 0.227 12 m3lh 1 m3lh 4.4029588 gaVrnin
1 gal/min 0.063088 Cld/S 1 elm% 15.850875 galirnin
lmi/h 88 ft/min 1 ftlmin 0.0113636 rnilh
lmi/h 1.609 lkmlh 0.621504 miIh

B.7 Specific Weights (Gases)


1 Ib/SCF dry = 16.882 kg/m3 dry lkg/m3dry = 0.0592347 IWSCF dry
1 1WSCF wet = 17.078 kg/m3 w e t 1 kg/m3 w e t = 0.0585549 IWSCF w e t
1 gr/sCJ? dry = 0.0024118 kg/m3 dry lkg/m3dry = 414.62808 gr/SCF dry
1 gr/SCF w e t = 0.0024397 kg/m3 w e t 1 kg/m3 wet = 409.88646 gr/SCJ! wet
SCF=30inHg@6OF 1.N. (normal) = 0 C, 1.01325 bar
B.8 Force

1 poundal = 0.138255 N 1N = 7.2330115 poundal


1 lb-force = 3.338221 N 1N = 0.2995608 lb-force
lkp a 9.80665 N 1N = 0.1019716 kp

B .9 Pressure
Note: Pa = N/m2

1 psi 6.8948 kNfm2 1 kNfm2 = 0.1450368 psi


1 psi 68.948 mbar 1 mbar = 0.0145037 psi
1 psi 2.3066 ft Hz0 1 ftH20 = 0.4335385 psi
1 psi 0.0703062 kg/cm2 1 kg/cm2 = 14.2235 psi
1 psi 703.07 kg/m2 1 kg/m2 = 0.00 14223 psi
1 in.Hg 0.038638 bar 1 bar = 25.881257 in. Hg
1 in.Hg 0.03342 atm 1 atm = 29.922202 in. Hg
1 in.Hg 3863.8 N/m2 1 N/m2 = 0.0002588 in. Hg
1 in. H20 0.002539 kg/cm2 1 kg/cm2 = 393.85585 in. H20
1 Iblft2 0.0004882 kg/cm2 1 kg/cm2 = 2048.3408 lb@
1 lblft2 47.876 N/m2 1 N/m2 = 0.0208873 lb/ft2
1 N/m2 0.000001 N/mm2 1 N/mm2 = 1OOOOOO N/m2
1 kp/cm2 0.0981 N/mm2 1 N/mm2 = 10.204082 kp/cm2
1 kp/cm2 98100 N/m2 1 N/m2 = 1.019E-05 kp/cm2
1N 0.102 kp lkp = 9.8039216 N
1 psi 6894.76 N/m2 1 N/m2 = 0.000145 psi
1 in. H20 24.899 N/m2 1 N/m2 = 0.0401623 in. Hz0
1 psi 0.0068953 N/mm2 1 N/mm2 = 145.02632 psi
13.1 1 Heat
1 Btu 0.2518892 kcal 1 kcal =: 3.96999 Btu
1 Btu 1.055056 kJ 1kJ = 0.9478 17 Btu
lBtu/lbF 1.00041 k&kg C 1 k&kg C = 0.9995902 Btu/lb F
1 BtuAb F 4.1886 kJ/kg c lkJ/kgC = 0.2387433 BtuIlb F
1 Btu/ft*hF 4.8844 kcaUm*hC 1 kcal/m2hC = 0.2047334 Btu/ft%F
1 Btu/ft%F 20.45 kJ/m*hC 1 kJ/mzhC = 0.0488998 Btu/ft*hF
1 Btu-inJft*h 0.068925 kcal/mh 1 kcavmh = 14.508524 Btu-inlftzh
1 Btu-inJft2h 0.288578 kJ/mh 1 kJ/mh = 3.4652676 BtwinJft2h
1 Btulfth 0.827 1 kcal/mh 1 kcal/mh. =: 1.2090436 Btu/ftb
1 Btu/fth 3.46294 kJ/mh 1 kJ/mh = 0.288772 Btu/fth
1 Btu-inJft*hF 0.12407 kCdhhC 1 kcavmhc =: 8.0599661 Btu-inlftzh
1 Btu-inlft%F 0.51946 kJ/ltlhC 1 kJ/mhc = 1.925076 Btu-inlftzh
1 Btdlb 0.55579 kcal/kg ll=avkg = 1‘7992407 Btu/lb
1 BNlb 2.327 kTkg 1 kJ/kg = 0.4297379 Btu/lb
1 Btu/ft* 2.7136 kcaUm2 1 kcal/m2 = 0.3685142 Btulftz
1 Btu/ftz 11.361138 kJlm* 1 M/m2 = 0.0880194 B tu/ft2
1 Btu/in? 390.76 kcal/m* 1 k&m2 = 0.0025591 Btu/in?
1 Bhdin.2 1636.1 kJlm2 1 kJlm2 = 0.0006112 BtuAn.2
1 Btdshort ton 0.000278 l=as ll=avkg E 3597.122 Btdhort ton
1 BNshort ton 0.0011631 kJ@ 1 M/kg = 859.77 13 Btu/short ton
1J 0.0002388 kcal lkd = 4187 J
1J 2.78E-07 kWh 1kWh = 36000 J
1 k&kg 4.184 AU/ton 1 MJ/ton =: 0.2390057 kcavks
1 kcaUm2h 1.163 W/m2 1 WIm* = 0.8598452 kcaVm2h
1 kcavmhc 1.163 W/mK 1 w/mK = 0.8598452 kcal/mhC
1 kcal/m*hC 1.163 W/Km2 1 w/Km2 = 0.8598452 kcal/m*hC
1 kcavkgc 4.184 kJ/kgK 1 kJ/kgK = 0.2390057 kcalAcgC
IS.11 Work and Energy
w = J/s
1 ft-lb (weight) = 1.35582 J 1J = 0.737561 ft-lb (weight)
1 ft-lb (force) = 1.35582 J 1J = 0.737561 ft-lb (force)
1 ft-poundal = 0.0421401 J 1J =: 23.730366 ft-poundal
1Btu = 1.055056 kJ 1kJ = 0.9478 17 Btu
1 Btu = 0.293071 Wh 1Wh t 3.4121425 Btu
1 hp-h = 2.68452 h-lJ 1MJ zz 0.3725061 M-h
1 hp-h = 0.7457 kWb 1kWh = 1.3410219 M-h
1 ft-lb (wt)/s = 1.35582 W 1w = 0.737561 ft-lb (wt)/s
1Btu z 1.055056 kW 1kW = 0.9478 17 Btu
1 Btu/h = 0.29307 kW 1kW = 3.4121541 Btu/h
1 hp = 0.7457 kW 1 kW = 1.3410219 hp
1 PS = 735.5 w 1w = 0.0013596 PS
1kWh =: 860 kcal 1 kcal = 0.0011628 kWh
1w =I 0.86 k&b 1 kcallh =: 1.1627907 W
1w = 0.001 kW 1kW = 1000 W
APPENDIX

B.12 Engineering Units

tera T trillion 1xE 12


giga G billion 1xE 9 1OOOOOOOOO
hundred million 1xE 8 1OOOOOOOO
ten million 1xE 7 1OOOOOOO
mega M million 1xE 6 ulomoo
hundred thousand 1xE 5 100000
ten thousand 1xE 4 loo00
kilo k thousand 1xE 3 1000
hecto h hundred 1xE 2 100
deca da ten 1xE 1 10
lx 1
deci d tenth lxE-1 0.1
centi c hundredth lxE-2 0.01
milli m thousandths lxE-3 0.001
ten thousandths lxE-4 o.ooo1
hundred thousandths lxE-5 O.ooool
micro P millionth lxE-6 o.oGQoo1
nano n billionth lxE-9 o.ooooooOO1
pica P trillionth 1 x E-12
THE ROTARY CEMENT KILN

C: TEMPERATURE CONVERSIONS
c F c F c
-34 -30 -22 -12 11 52 11 52 126
-34 -29 -20 -11 12 54 12 53 127
-33 -28 -18 -11 13 55 12 54 129
-33 -27 -17 -10 14 57 13 55 131
-32 -26 -15 -9 15 59 13 56 133
-32 -25 -13 -9 16 61 14 57 135
-31 -24 -11 -8 17 63 14 58 136
-31 -23 -9 -8 18 64 15 59 138
-30 -22 -8 -7 19 66 16 60 140
-29 -21 -6 -7 20 68 16 61 142
-29 -20 4 -6 21 70 17 62 144
-28 -19 -2 4 22 72 17 63 145
-28 -18 0 -5 23 73 18 64 147
-27 -17 1 4 24 75 18 65 149
-27 -16 3 4 25 77 19 66 151
-26 -15 5 -3 26 79 19 67 153
-26 -14 7 -3 27 81 20 68 154
-25 -13 9 -2 28 82 21 69 156
-24 -12 10 -2 29 84 21 70 158
-24 -11 12 -1 30 86 22 71 160
-23 -10 14 -1 31 88 22 72 162
-23 -9 16 0 32 90 23 73 163
-22 4 18 1 33 91 23 74 165
-22 -7 19 1 34 93 24 75 167
-21 -6 21 2 35 95 24 76 169
-21 -5 23 2 36 97 25 77 171
-20 -4 25 3 37 99 26 78 172
-19 -3 27 3 38 100 26 79 174
-19 -2 28 4 39 102 27 80 176
-18 -1 30 4 40 104 27 81 178
-18 0 32 5 41 106 28 82 180
-17 1 34 6 42 108 28 83 181
-17 2 36 6 43 109 29 84 183
-16 3 37 7 44 111 29 85 185
-16 4 39 7 45 113 30 86 187
-15 5 41 8 46 115 31 87 189
-14 6 43 8 47 117 31 88 19c
-14 7 45 9 48 118 32 89 192
-13 8 46 9 49 120 32 90 194
-13 9 48 10 50 122 33 91 1%
-12 10 50 11 51 124 33 92 198
APPENDIX

C F C F C F

34 93 199 57 135 275 81 177 351


34 94 201 58 136 277 81 178 352
35 95 203 58 137 279 82 179 354
36 96 205 59 138 280 82 180 356
36 97 207 59 139 282 83 181 358
37 98 208 60 140 284 83 182 360
37 99 210 61 141 286 84 183 361
38 100 212 61 142 288 84 184 363
38 101 214 62 143 289 85 185 365
39 102 216 62 144 291 86 186 367
39 103 217 63 145 293 86 187 369
40 104 219 63 146 295 87 188 370
41 105 221 64 147 297 87 189 372
41 106 223 64 148 298 88 190 374
42 107 225 65 149 300 88 191 376
42 108 226 66 150 302 89 192 378
43 109 228 66 151 304 89 193 379
43 110 230 67 152 306 !xl 194 381
44 111 232 67 153 307 91 195 383
44 112 234 68 154 309 91 196 385
45 113 235 68 155 311 92 197 387
46 114 237 69 156 313 92 198 388
46 115 239 69 157 315 93 199 390
47 116 241 70 158 316 93 200 392
47 117 243 71 159 318 94 201 394
48 118 244 71 160 320 94 202 3%
48 119 246 72 161 322 95 203 397
49 120 248 72 162 324 % 204 399
49 121 250 73 163 325 % 205 401
50 122 252 73 164 327 97 206 403
51 123 253 74 165 329 97 207 405
51 124 255 74 166 331 98 208 406
52 125 257 75 167 333 98 209 408
52 126 259 76 168 334 99 210 410
53 127 261 76 169 336 99 211 412
53 128 262 77 170 338 100 212 414
54 129 264 77 171 340 101 213 415
54 130 266 78 172 342 101 214 417
55 131 268 78 173 343 102 215 419
56 132 270 79 174 345 102 216 421
56 133 271 79 175 347 103 217 423
57 134 273 80 176 349 103 218 424
THE ROTARY CEMENT KILN

C F C F C F

104 219 426 152 305 581 268 515 959


104 220 428 154 310 590 271 520 968
105 221 430 157 315 599 274 525 977
106 222 432 160 320 608 277 530 986
106 223 433 163 325 617 279 535 995
107 224 435 166 330 626 282 540 1004
107 225 437 168 335 635 285 545 1013
108 226 439 171 340 644 288 550 1022
108 227 441 174 345 653 291 555 1031
109 228 442 177 350 662 293 560 1040
109 229 444 179 355 761 2% 565 1049
110 230 446 182 360 680 299 570 1058
111 231 448 185 365 689 302 575 1067
111 232 450 188 370 698 304 580 1076
112 233 451 191 375 707 307 585 1085
112 234 453 193 380 716 310 590 1094
113 235 455 1% 385 725 313 595 1103
113 236 457 199 390 734 316 600 1112
114 237 459 202 395 743 318 605 1121
114 238 460 204 400 752 321 610 1130
115 239 462 207 405 761 32A 615 1139
116 240 464 210 410 770 327 620 1148
116 241 466 213 415 779 329 625 1157
117 242 468 216 420 788 332 630 1166
117 243 469 218 425 797 335 635 1175
118 244 471 221 430 806 338 640 1184
118 245 473 224 435 815 341 645 1193
119 246 475 227 440 824 343 650 1202
119 247 477 229 445 833 349 660 1220
120 248 478 232 450 842 354 670 1238
121 249 480 235 455 851 360 680 1256
121 250 482 238 460 860 366 690 1274
124 255 491 241 465 869 371 700 1292
127 260 500 243 470 878 377 710 1310
129 265 509 246 475 887 382 720 1328
132 270 518 249 480 8% 388 730 1346
135 275 527 252 485 905 393 740 1364
138 280 536 254 490 914 399 750 1382
141 285 545 257 495 923 404 760 1400
143 290 554 260 500 932 410 770 1418
146 295 563 263 505 941 416 780 1436
149 300 572 266 510 950 421 790 1454
APPENDIX

C F C F C F

427 800 1472 660 1220 2228 893 1640 2984


432 810 1490 666 1230 2246 899 1650 3002
438 820 1508 671 1240 2264 904 1660 3020
443 830 1526 677 1250 2282 910 1670 3038
449 840 1544 682 1260 2300 916 1680 3056
454 850 1562 688 1270 2318 921 1690 3074
460 860 1580 693 1280 2336 927 1700 3092
466 870 1598 699 1290 2354 932 1710 3110
471 880 1616 704 1300 2372 938 1720 3128
477 890 1634 710 1310 2390 943 1730 3146
482 900 1652 716 1320 2408 949 1740 3164
488 910 1670 721 1330 2426 954 1750 3182
493 920 1688 727 1340 2444 960 1760 3200
499 930 1706 732 1350 2462 966 1770 3218
504 940 1724 738 1360 2480 971 1780 3236
510 950 1742 743 1370 2498 977 1790 3254
516 960 1760 749 1380 2516 982 1800 3272
521 970 1778 754 1390 2534 988 1810 3290
527 980 17% 760 1400 2552 993 1820 3308
532 990 1814 766 1410 2570 999 1830 3326
538 1000 1832 771 1420 2588 1004 1840 3344
543 1010 1850 777 1430 2606 1010 1850 3362
549 1020 1868 782 1440 2624 1021 1870 3398
554 1030 1886 788 1450 2642 1032 1890 3434
560 1040 1904 793 1460 2660 1043 1910 3470
566 1050 1922 799 1470 2678 1054 1930 3506
571 1060 1940 804 1480 26% 1066 1950 3542
577 1070 1958 810 1490 2714 1077 1970 3578
582 1080 1976 816 1500 2732 1088 1990 3614
588 1090 1994 821 1510 2750 1099 2010 3650
593 1100 2012 827 1520 2768 1110 2030 3686
599 1110 2030 832 1530 2786 1121 2050 3722
604 1120 2048 838 1540 2804 1132 2070 3758
610 1130 2066 843 1550 2822 1143 2090 3794
616 1140 2084 849 1560 2840 1154 2110 3830
621 1150 2102 854 1570 2858 1166 2130 3866
627 1160 2120 860 1580 2876 1177 2150 3902
632 1170 2138 866 1590 2894 1188 2170 3938
638 1180 2156 871 1600 2912 1199 2190 3974
643 1190 2174 877 1610 2930 1210 2210 4oia
649 1200 2192 882 1620 2948 1221 2230 4046
654 1210 2210 888 1630 2966 1232 2250 4082
382 THE ROTARY CEMENT KILN

C F C F C F

1243 2270 4118 1388 2.530 4586 1532 2790 5054


1254 2290 4154 1399 2550 4622 1543 2810 5090
1266 2310 4190 1410 2570 4658 1554 2830 5126
1277 2330 4226 1421 2590 4694 1566 2850 5162
1288 2350 4262 1432 2610 4730 1577 2870 5198
1299 2370 4298 1443 2630 4766 1588 2890 5234
1310 2390 4334 1454 2650 4802 1599 2910 5270
1321 2410 4370 1466 2670 4838 1610 2930 5306
1332 2430 4406 1477 2690 4874 1621 2950 5342
1343 2450 4442 1488 2710 4910 1632 2970 5378
1354 2470 4478 1499 2730 4946 1643 2990 5414
1366 2490 4514 1510 2750 4982 1654 3010 5450
1377 2510 4550 1521 2770 5018 1666 3030 5486
c

D: KILN OPERATOR’S QUIZ


This quiz is &signed to give the reader an opportunity to evaluate his
knowledge of kiln operation. The quiz consists of 50 true-or-false
questions .

D.1 FUNDAMENTALS (Score 5 points for each correct answer)

T F
1. When air is heated, its weight per unit volume de- - -
creases.
2. One ton of kiln feed will make one ton of clinker. - -
3. High free-lime contents in the clinker always indicate - -
that the clinker has been underburned.
4. C$ (tricalcium silicate) is an ingredient in the kiln - -
feed.
5. CO2 (carbon dioxide) in the kiIn exit gases originates - -
partly from the combustion of the fuel and partly
from calcination of feed.
6. An increase in fuel rate will always generate more ____ -
heat in the burning zone.
7. Opening a butterfly damper from 80% to 100% open - -
will most likely produce no increase in flow.
8. Iron and alumina are fluxing agents in the kiln feed
mix and tend to make the clinker easier to bum.
9. On coal-fired kilns, the chemical composition of the
clinker will be the same as the composition of the
kiln feed.
10. Large changes in the silica ratio will indicate possible
changes in the bumability of the clinker.
11. A lime saturation factor of 1.03 and a free lime of
1.4% in the clinker means the clinker has been
underburned.
12. The predominant component in the kiln feed is
calcium carbonate.
13. Calcium oxide (CaO) is the product when limestone
has been calcined.
14. All metal components of the kiln, when heated, will
contract.
15. Higher undergrate pressures in the cooler compart-
ments indicate a higher air-flow rate.
16. If the fan speed and damper setting remain the same,
the amperes on an air fan will increase when the air
temperature decreases.
17. The oxygen content in the kiln exit gases is a sole
function of the amount of air passing through the
kiln.
18. Red color on the kiln shell means there is no refrac-
tory lining left.
19. Higher bed depth in the cooler produces better cooling
of the clinker.
20. A higher kiln speed shortens the residence time of the
feed in the kiln.

DJ OPERATIONAL
(Note: Assume all other kiln variables remain constant)
(Score 2 pints for each correct answer)
T F
21. An increase in fuel rate will cause a decrease in the - -
kiln exit-gas oxygen content.
22. Slight percentages of combustibles in the kiln exit - -
gas can be ignored.
384 THE ROTARY CEMENT KILN

23. Ideally, a kiln operates with approximately 5% excess


air.
24. In normal operation, the kiln exit gas should show
approximately 3% oxygen.
25. An increase in I.D. fan speed will increase the
burning-zone temperature.
2 6 . The burning-zone temperature can only be increased
by increasing the fuel rate.
27. A decrease in I.D. fan speed will result in a decrease
in burning-zone temperature.
28. When a ring is building up in the kiln, the feed-end
draft will increase.
29. The amount of air entering the kiln proper is
governed by the amount of air that is forced into the
cooler.
3 0 . Opening the cooler stack damper means less air will
enter the kiln proper.
3 1 . A positive hood pressure is solely caused by too
much air being forced into the cooler.
32. A drop in kiln speed will cause the feed-end
temperature to rise.
33. A reduction in feed rate will cause the feed-end
temperature to fall.
3 4 . One ton of kiln dust returned will produce the same
amount of clinker as one ton of fresh kiln feed.
3 5 . It is generally easier to form coating on coal- than gas-
fired kilns.
3 6 . Changing from natural gas to coal firing will not
alter the chemical composition and bumability of the
resultant clinker.
37. Increasing the cooler-grate drive speed will bring
about a thinner clinker bed.
38. An increase in the cooler exhaust-fan speed (or
opening the damper at the cooler exhaust stack) will
cause a decrease in the hood pressure.
3 9 . A kiln will always operate better at low production
rates than at high rates.
40. An increase in the feed ratio will result in higher
production rates.
APPENDIX

41. Frequent kiln stops and upsets tend to shorten the - -


refractory life.
42. A rotary kiln is never perfectly round. - -
43. At the beginning of a shift, the operator should - __
always change the controller settings to confrom to
the settings as they were 24 h earlier when the kiln
operated in stable fashion.
44. An increase in feed rate will result in a lower feed-end - -
temperature.
45. Clinker, when burned at a higher temperature, wilI
show a higher density.
46. When the oxygen in the exit gases shows 0.4% and
the kiln is cooling down, the operator should increase
the fuel rate.
47. When starting a kiln, the primary-air and I.D. fans - -
must always be started first before fuel is given to the
kiln.
48. One should always try first to secure stable operations _ _ -
before an attempt is made to push the kiln to
maximum production.
49. Too much air flow into the cooler can impair clinker _ _ ~
advancement through the cooler.
50. The only way one can increase the production rate of - _ _
the kiln is to increase the kiln speed.

ANSWERS TO QZJZZ
1. T 11. F 21. T 31. F 41. T
2. F 12. T 22. F 32. T 42. T
3. F 13. T 23. T 33. F 43. F
4. F 14. F 24. F 34. F 44. T
5. T 15. F 25. F 35. T 45. T
6. F 16. T 26. F 36. F 46. F
7. T 17. F 27. F 37. T 47. T
8. T 18. F 28. T 38. T 48. T
9. F 19. F 29. F 39. F 49. T
I 10. T 20. T 30. F 40. T 50. F
Index

A C
Accidents, 363 COz, 58,73, 166, 167,246,293
Air, circuit, 155, 158, 165 Calcination, 85, 88
deficiency, 53 percent, 60, 88
distribution, cooler, 283 Calcining zone, 59, 142
for combustion, 50, 166 Capacity, of kiln, 179, 197
inleakage, 112, 173 Carbon dioxide, 58
pressure, 155 monoxide, 52
velocity, 155 Cardox system, 152
Alite, 139 Cement composition, 121
Alkalies, 117 Chains, 8
Alumina bricks, 29 faihlre, 20
Alumina-Iron ratio, 125 moisture, 144
Ash composition, 119 Chlorides, 118
Automatic control, 206 Clinker, composition, 121, 122,
137
B compounds, 85
microscopy, 138
Back-end temperature, 239, 246, properties, 131
292,327 rings, 150
Baghouse, 172 size, 274
Basic refractory, 29, 151 temperature, 112,235, 289
Belite, 139 coal, 37
Birefringence, 140 ash, 119, 120
Black feed, 232,301 factor, 49
Bogue formulas, 121 grinding, 73,254
Brick charts, 35 mill temperatures. 72
Bricks required, 33 storage, 38, 254
Btu’s 42 Coating, 88, 147,238
Bumability, 124, 128 Combustibles, 53
Burner alignment, 79 Combustion, 44, 50, 60, 166,240
tip velocity, 65,260 air, 50.57, 66
Burning technique, 69,327 control, 56,68,71
Burning zone, 143, 152 products, 58
control, 232,327 precalciner, 57
temperature, 235,239 Computer control, 230
INDEX

Conduction, 83 direction, 77
c o n t r o l 1OODS. 229 ignition, 69
Controllers,- 2b7.228 length, 64
convection. 83 propagation, 66
Conversion’factors, 367 shape, 77, 236
Cooler, air, 158, 166, 283 temperature, 81
bed depth, 279 Flames, 63,259
clinker distribution, 283 Forced fining, 179
control, 267 Fuel, burners, 259
fans, 280 coal 37,49
planetary, 162 control, 244
reciprocating, 161, 276 Et=, 41
retention time, 273 grinding, 253
rotary, 163 handling, 253
stack losses, 112 oil, 40
Cooling zone, 143
Cycle, 320 G
D G-Cooler, 161
Gas, analyzer, 59
Dark feed, 235 flow rate, 167-171
Dicalcium silicate, 123 laws, 44
Direct coal firing, 39 losses, 108
Dolomite refractory, 30, 150 pressures, 45
Draft, 166,294, 297 temperatures, 46
Dry process, 8, 110, 145 velocity, 157, 180
Drying zone, 142 volumes, 44
Dust collectors, 172 Grate coolers, 161,276
Dust losses, 108
H
E
Hazards, 364
Emergency conditions, 334, 346 Heat, balance, 106
Exit gas, 240, 272, See also back- exchangers, 86
end. profile 84, 88
savings, 113
F transfer, 83, 94, 164
Heating value, 42
Fan, damper position, 282 High-grade heat, 85
horsepower, 281 Hood draft, 290
static pressure, 281 Hydraulic ratio, 126
Fans, 280
Feed, 22 I
Feed end temperature, See back-
end. I.D. fan, 20, 66, 69, 172, 245,
Feed rate, 299,302 290.295
Feed, behind burning zone, 235 IS0 shapes, 31,33, 34
Flame, adjustment, 78 Ignition, 69
THE ROTARY CEMENT KILN

Indirect coal ftig, 39 Nose rings, 150


Inleakage, of air, 173 NO, 243,246
Instrumentation, 207,213
Interlocks, 229 0
K Ono microscopic test, 138
Operating methods, 201,327-345
Kiln, 27 conditions, 327 Oxygen, 53, 69, 239, 241, 244,
capacity, 179 327
co&~ 237 enrichment, 75
cycling, 320
diameter, 35 P
draft, 294
drive amperes, 239 PPM, 54
drive torque, 239,246 Petroleum coke, 37
feed, 115, 120 Phase formations, 85
feed fineness, 130 Planetary coolers, 162
loading, 174 Plume, 70
output, 300 Potassium, 117
roll back, 250 Precalciner, air circuif 160
shell, 3 12 kh, 5, lo,75 87, 146
shutdowns, 71,313 Preheat zone, 142
speed, 86,248,306 Preheater, air circuit, 159
start, 69,308 kiln, 10,86, 111, 146
tires, 312 Pressure measurements, 156
turning, 20,318 F’rimq air, 67,70
fan, 20
L temperature, 256
volume, 261
Lepol kilns, 7,146
Lignite, 37
Lime saturation factor, 126 R
Liquid formed, 150
Liquid, percent, 127 Radiation, 83
Liter weight test, 135 Ransome kiln, 4
Low grade heat, 85 Raw materials, 115, 116
Loading of kiln, 175, 177 Recorders, 207,209
Refractory, 17,28,309
M failure, 17
life, 19
Magnesia, 116 properties, 22-27
Microscope, 138 shapes, 18, 31.33
Mudrings, 154 DIN tests, 27
Movement of kihr feed, 171 heat transfer, 94
Retention time, 176,
N 305
Ring formation, 147, 153
Natural gas, 41 Rotary coolers, 163
INDEX

S Tire, expansion, 312


slippage, 22
Safety, 362 Transition zones, 143
se-h-y air, 70,237,287 Travel time, See retention time.
Semidry process, 7 Tricalcium aluminate, 123
Shaft kilns, 3 silicate, 123
Shell, expansion, 21, 312
oval@, 22 U
temperatures, 100, 112
Silica ratio, 125 Undergrate pressure, 267, 279
Sinter zone, 143
Slurry, 144
moisture, 108 V
Sodium, 117
Solid fuels, 37 Velocity, gas, 157, 180
Spinell bonded bricks, 31 vertical kilns. 3
Stability, operation, 203 Viewing glass, 233
sulfur, 117 Volatiles. 67
System check, 309 Volumes changes, refractory, 3 1 I
Volume, gas, 157
T
W
Technique, burning, 201
Temperature conversion, 378 Wet-process kilns, 6, 109, 113,
Tertiary air, 12 145
Thermal, contraction, 317
expansion, 3 11 z
work in kiln, 84
Tip velocity, 260 Zones in kiln, 141

Potrebbero piacerti anche